This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

\AtEndEnvironment

proof

Definable convolution and idempotent Keisler measures III. Generic stability, generic transitivity, and revised Newelski’s conjecture

Artem Chernikov Artem Chernikov [Uncaptioned image]https://orcid.org/0000-0002-9136-8737
University of Maryland, College Park and University of California, Los Angeles, USA
[email protected]
Kyle Gannon Kyle Gannon
Beijing International Center for Mathematical Research (BICMR)
Peking University
Beijing, China
[email protected]
 and  Krzysztof Krupiński Krzysztof Krupiński [Uncaptioned image]https://orcid.org/0000-0002-2243-4411
Instytut Matematyczny Uniwersytet Wrocławski, pl. Grunwaldzki 2, 50-384 Wrocław, Poland
[email protected]
Abstract.

We study idempotent measures and the structure of the convolution semigroups of measures over definable groups.

We isolate the property of generic transitivity and demonstrate that it is sufficient (and necessary) to develop stable group theory localizing on a generically stable type, including invariant stratified ranks and connected components. We establish generic transitivity of generically stable idempotent types in important new cases, including abelian groups in arbitrary theories and arbitrary groups in rosy theories, and characterize them as generics of connected type-definable subgroups.

Using tools from Keisler’s randomization theory, we generalize some of these results from types to generically stable Keisler measures, and classify idempotent generically stable measures in abelian groups as (unique) translation-invariant measures on type-definable fsg subgroups. This provides a partial definable counterpart to the classical work of Rudin, Cohen and Pym for locally compact topological groups.

Finally, we provide an explicit construction of a minimal left ideal in the convolution semigroup of measures for an arbitrary countable NIP group, from a minimal left ideal in the corresponding semigroup on types and a canonical measure constructed on its ideal subgroup. In order to achieve it, we in particular prove the revised Ellis group conjecture of Newelski for countable NIP groups.

1. Introduction

We study idempotent measures and the structure of the convolution semigroups on measures in definable groups, as well as some related questions about topological dynamics of definable actions (continuing [CG22, CG23]).

We first recall the classical setting. If GG is a locally compact group and (G)\mathcal{M}(G) is the space of regular Borel probability measures on GG, one extends group multiplication on GG to convolution \ast on (G)\mathcal{M}(G): if μ,ν(G)\mu,\nu\in\mathcal{M}(G) and BB is a Borel subset of GG, then

(μν)(B)=GG𝟏B(xy)𝑑μ(x)𝑑μ(y).(\mu*\nu)(B)=\int_{G}\int_{G}\mathbf{1}_{B}(x\cdot y)d\mu(x)d\mu(y).

A measure μ\mu is idempotent if μμ=μ\mu*\mu=\mu. A classical line of work established a correspondence between compact subgroups of GG and idempotent measures in (G)\mathcal{M}(G), in progressively broader contexts [KI40, Wen54, Coh60, Rud59, Gli59] culminating in the following:

Fact 1.1.

[Pym62, Theorem A.4.1] Let GG be a locally compact group and μ(G)\mu\in\mathcal{M}(G). Then the following are equivalent:

  1. (1)

    μ\mu is idempotent.

  2. (2)

    The support supp(μ)\operatorname{supp}(\mu) of μ\mu is a compact subgroup of GG and μ|supp(μ)\mu|_{\operatorname{supp}(\mu)} is the normalized Haar measure on supp(μ)\operatorname{supp}(\mu).

We are interested in a counterpart of this phenomenon in the definable category. In the same way as e.g. algebraic or Lie groups are important in algebraic or differential geometry, the understanding of groups definable in a given first-order structure (or in certain classes of first-order structures) is important for model theory and its applications. The class of stable groups is at the core of model theory, and the corresponding theory was developed in the 1970s-1980s borrowing many ideas from the study of algebraic groups over algebraically closed fields (with corresponding notions of connected components, stabilizers, generics, etc., see [Poi01]). More recently, many of the ideas of stable group theory were extended to the class of NIP groups, which contains both stable groups and groups definable in oo-minimal structures or over the pp-adics. This led to multiple applications, e.g. a resolution of Pillay’s conjecture for compact o-minimal groups [HPP08] or Hrushovski’s work on approximate subgroups [Hru12], and brought to light the importance of the study of invariant measures on definable subsets of the group (see e.g. [Che18] for a short survey), as well as the methods of topological dynamics (introduced into the picture starting with Newelski [New09]). In particular, deep connections with tame dynamical systems as studied by Glasner, Megrelishvili and others (see e.g. [Gla07, Gla18]) have emerged, and play an important role in the current paper.

More precisely, we now let GG be a group definable in some structure MM (i.e. both the underlying set and multiplication are definable by formulas with parameters in MM), it comes equipped with a collection of definable subsets of cartesian powers of GG closed under Boolean combinations, projection and Cartesian products (but does not carry topology or any additional structure a priori). We let 𝒰\mathcal{U} be a “non-standard” elementary extension of MM, and we let G(𝒰)G(\mathcal{U}) denote the group obtained by evaluating in 𝒰\mathcal{U} the formulas used to define GG in MM (which in the case of an algebraic group corresponds to working in the universal domain, in the sense of Weil). So e.g. if we start with M=(,+,×)M=(\mathbb{R},+,\times) the field of reals, and G(M)G(M) its additive group, then G(𝒰)G(\mathcal{U}) is the additive group of a large real closed field extending \mathbb{R} which now contains infinitesimals — i.e., it satisfies a saturation condition: every small finitely consistent family of definable sets has non-empty intersection. It is classical in topological dynamics to consider the action of a discrete group GG on the compact space βG\beta G of ultrafilters on GG, or more precisely ultrafilters on the Boolean algebra of all subsets of GG. In the definable setting, given a definable group G(M)G(M), we let SG(M)S_{G}(M) denote the space of ultrafilters on the Boolean algebra of definable subsets of G(M)G(M), hence the space SG(M)S_{G}(M) (called the space of types of G(M)G(M)) is a “tame” analogue of the Stone-Čech compactification of the discrete group GG. Then G(M)G(M) acts on SG(M)S_{G}(M) by homeomorphisms, and the same construction applies to G(𝒰)G(\mathcal{U}) giving the space SG(𝒰)S_{G}(\mathcal{U}) of ultrafilters on the definable subsets of G(𝒰)G(\mathcal{U}). Similarly, we let 𝔐G(M)\mathfrak{M}_{G}(M) denote the space of finitely additive probability measures on the Boolean algebra of definable subsets of G(M)G(M) (and 𝔐G(𝒰)\mathfrak{M}_{G}(\mathcal{U}) for G(𝒰)G(\mathcal{U})), it is affinely homeomorphic to the space of all regular σ\sigma-additive Borel probability measures on SG(M)S_{G}(M) (respectively on SG(𝒰)S_{G}(\mathcal{U})), with weak-topology. The set G(M)G(M) embeds into SG(𝒰)S_{G}(\mathcal{U}) as realized types, and we let SG,M(𝒰)S_{G,M}(\mathcal{U}) denote its closure (model theoretically, this corresponds to the set of global types in SG(𝒰)S_{G}(\mathcal{U}) that are finitely satisfiable in G(M)G(M)). Similarly, we let 𝔐G,M(𝒰)\mathfrak{M}_{G,M}(\mathcal{U}) denote the closed convex hull of G(M)G(M) in 𝔐G(𝒰)\mathfrak{M}_{G}(\mathcal{U}) (this is the set of global Keisler measures on G(𝒰)G(\mathcal{U}) finitely satisfiable in G(M)G(M), equivalently the set of measures supported on SG,M(𝒰)S_{G,M}(\mathcal{U}) — see [CG22, Proposition 2.11]). Similarly to the classical case, in many situations (including the ones discussed in the introduction) we have a well-defined convolution operation \ast on 𝔐G,M(𝒰)\mathfrak{M}_{G,M}(\mathcal{U}) (see Definition 3.35 and the discussion around it).

In this context, generalizing a classical fact about idempotent types in stable groups [New91], we have the following definable counterpart of Fact 1.1 for stable groups:

Fact 1.2.

[CG22, Theorem 5.8] Let GG be a (type-)definable group in a stable structure MM and μ𝔐G,M(𝒰)\mu\in\mathfrak{M}_{G,M}(\mathcal{U}) a measure. Then μ\mu is idempotent if and only if μ\mu is the unique left-invariant (and the unique right-invariant) measure on a type-definable subgroup of G(𝒰)G(\mathcal{U}) (namely, the left-/right-stabilizer of μ\mu).

This suggests a remarkable analogy between the topological and definable settings, even though Fact 1.2 is proved using rather different methods.

In the first part of the paper (Sections 2 and 3), we study generalizations of Fact 1.2 beyond the limited context of stable groups (we note that this correspondence fails in general NIP groups without an appropriate tameness assumption on the idempotent measure [CG23, Example 4.5]). An important class of groups arising in the work on Pillay’s conjectures is that of groups with finitely satisfiable generics, or fsg groups in short [HPP08]. It contains stable groups, as well as (definably) compact groups in oo-minimal structures, and provides a natural counterpart to the role that compact groups play in Fact 1.2. By a well-known characterization in the NIP context (see e.g. [Sim15, Proposition 8.33]), these are precisely the groups that admit a (unique) translation-invariant measure μ\mu on their definable subsets which is moreover generically stable: a sufficiently long random sample of elements from the group uniformly approximates the measure of all sets in a definable family of subsets with high probability (i.e. μ\mu is a frequency interpretation measure, or fim measure, satisfying a uniform version of the weak law of large numbers — this notion is motivated by Vapnik-Chervonenkis theory, and serves as a correct generalization of generic stability for measures outside of NIP, by analogy with generically stable types in the sense of [PT11]); see Section 3.3). An analog of Fact 1.2 would thus amount to demonstrating that such subgroups are the only source of idempotent generically stable measures (see Problem 3.41).

First, in Section 2 we focus on the case of idempotent types in SG,M(𝒰)S_{G,M}(\mathcal{U}) (i.e. {0,1}\{0,1\}-measures, equivalently ultrafilters on the Boolean algebra of definable subsets of GG). After reviewing some preliminaries on generically stable types (Sections 2.1 and 2.4), we revise the case of groups in stable structures (Section 2.7), and then resolve the question in several important cases:

Theorem 1.3.

Assume pSG(𝒰)p\in S_{G}(\mathcal{U}) is generically stable and idempotent, and one of the following holds:

  1. (1)

    pp is stable and MM is arbitrary (Proposition 2.29, see Section 2.8);

  2. (2)

    GG is abelian and MM is arbitrary (Proposition 2.18, see Section 2.5);

  3. (3)

    GG is arbitrary and MM is inp-minimal (Proposition 2.23, see Section 2.6);

  4. (4)

    GG is arbitrary and MM is rosy (so e.g. if MM has a simple theory; Proposition 2.35, see Sections 2.10 and 2.9).

Then pp is the unique left-/right-invariant type on a type-definable subgroup of G(𝒰)G(\mathcal{U}) (namely, the left-/right-stabilizer of pp).

The proof proceeds by establishing the crucial property of generic transitivity (see Section 2.4) for idempotent generically stable types in these cases, namely that if (a1,a2)pp(a_{1},a_{2})\models p\otimes p, then (a1a2,a1)pp(a_{1}\cdot a_{2},a_{1})\models p\otimes p (using local weight arguments in case (2), and the appropriate version of the theory of stratified ranks in the other cases). The question whether every generically stable idempotent type is generically transitive remains open, even for NIP groups (see Problem 2.15 and discussion in Section 2.4).

We further investigate generic transitivity, and demonstrate that it is a sufficient and necessary condition for developing some crucial results of stable group theory localizing on a generically stable type (some other elements of stable group theory for generically stable types were considered in [Wan22]). Sometimes we use a slightly stronger technical assumption that p(n)p^{(n)} is generically stable for all nn, which always holds in NIP structures. In Section 2.11, working in an arbitrary theory, we define an analog of the stratified rank in stable theories restricting to subsets of G(𝒰)G(\mathcal{U}) defined using parameters from a Morley sequence in a generically stable type pp, demonstrate finiteness of this rank (Lemma 2.39) and show that this rank is left invariant (under multiplication by realizations of pp) if and only if pp is generically transitive (Proposition 2.41). A fundamental theorem of Hrushovski [Hru90] demonstrates that in a stable theory, every type definable group (i.e. an intersection of definable sets that happens to be a group) is in fact an intersection of definable groups. The main result of Section 2.12 is an analog for generically transitive types:

Theorem 1.4 (Proposition 2.44).

If GG is type-definable and pSG(𝒰)p\in S_{G}(\mathcal{U}) is generically stable, idempotent and generically transitive, then its stabilizer is an intersection of MM-definable groups.

Finally, in Section 2.13 we establish a chain condition for groups type-definable using parameters from a Morley sequence of a generically stable type pp, implying that there is a smallest group of this form — and it is equal to the stabilizer of pp when pp is generically transitive (see Lemma 2.46 and Proposition 2.50 for the precise statement).

In Section 3, we generalize some of these results from types (i.e. {0,1}\{0,1\}-measures) to general measures, in arbitrary structures. Our main result is a definable counterpart of Fact 1.1 for arbitrary abelian group:

Theorem 1.5.

(Theorem 3.45) Let GG be an abelian group and μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) a generically stable measure. Then μ\mu is idempotent if and only if μ\mu is the unique left-invariant (and the unique right-invariant) measure on a type-definable subgroup of G(𝒰)G(\mathcal{U}) (namely, its stabilizer).

Groups as in Theorem 1.5, i.e. supporting an invariant generically stable measure, are called fim groups (see Section 3.6), and in the NIP case correspond precisely to fsg groups (but this is potentially a stronger condition in general). Our proof of Theorem 1.5 relies on several ingredients of independent interest. First, we develop some theory of fim  groups, generalizing from fsg groups in NIP structures (Section 3.6). Then, in Section 3.4, we provide a characterization of generically stable measures of independent interest extending [CGH23], demonstrating that the usual property — any Morley sequence determines the measure of arbitrary formulas by averaging along it — holds even when the parameters of these formulas are allowed to be “random”. More precisely:

Theorem 1.6 (Theorem 3.13).

Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) be fim, ν𝔐y(𝒰)\nu\in\mathfrak{M}_{y}(\mathcal{U}) arbitrary, φ(x,y,z)\varphi(x,y,z) a formula, b𝒰zb\in\mathcal{U}^{z}, and let 𝐱=(xi)iω\mathbf{x}=(x_{i})_{i\in\omega}. Suppose that λ𝔐𝐱y(𝒰)\lambda\in\mathfrak{M}_{\mathbf{x}y}(\mathcal{U}) is arbitrary such that λ|𝐱,M=μ(ω)\lambda|_{\mathbf{x},M}=\mu^{(\omega)} and λ|y=ν\lambda|_{y}=\nu. Then

limiλ(φ(xi,y,b))=μν(φ(x,y,b)).\lim_{i\to\infty}\lambda(\varphi(x_{i},y,b))=\mu\otimes\nu(\varphi(x,y,b)).

Moreover for every ε>0\varepsilon>0 there exists n=n(μ,φ,ε)n=n(\mu,\varphi,\varepsilon)\in\mathbb{N} so that for any ν,λ,b\nu,\lambda,b as above, we have λ(φ(xi,y,b))εμν(φ(x,y,b))\lambda(\varphi(x_{i},y,b))\approx^{\varepsilon}\mu\otimes\nu(\varphi(x,y,b)) for all but nn many iωi\in\omega.

This is new even in the NIP case, and relies on the use of Keisler randomization theory. Namely, we use the correspondence between measures in G(𝒰)G(\mathcal{U}) and types in its randomization, viewed as a structure in continuous logic, that was introduced in [BY09] (and studied further in [CGH23]). It allows us to imitate in Section 3.8 the bounded local weight argument from Section 2.5 in a purely measure theoretic context, using an adapted version of generic transitivity (see Section 3.7) and arguments with pushforwards.

Problem 2.15 on whether every generically stable idempotent type is generically transitive generalizes to measures (see Problem 3.41). In Section 3.9, we distinguish a weaker property of a measure than being generically transitive, which we call support transitivity. It leads to a weaker conjecture saying that every generically stable idempotent measure is support transitive (see Problem 3.48). While this conjecture is open, it trivially holds for idempotent types, and so one can expect that if the techniques used for types in Sections 2.72.10 could be adapted to measures, they would rather not prove the main conjecture that every generically stable idempotent measure is generically transitive, but reduce it to the above weakening. The idea is to pass to the randomization of the structure in question, and if this randomization happens to have a well-behaved stratified rank, then apply a continuous logic version of the arguments from Sections 2.72.10. In Section 3.10, we illustrate how it works for stable theories (recall that stability is preserved under randomization).

In Section 4, instead of considering an individual (idempotent) measure, we study the structure of the (left-continuous compact Hausdorff) semigroup (𝔐G,M(𝒰),)\left(\mathfrak{M}_{G,M}(\mathcal{U}),\ast\right) of measures on a definable NIP group under convolution, through the lens of Ellis theory. It was demonstrated in [CG23, Theorem 5.1] that the ideal (or Ellis) subgroup of any minimal left ideal is always trivial, and that when GG is definably amenable (i.e. admits a left-invariant finitely additive probability measure on its definable subsets), then any minimal left ideal itself is trivial, but has infinitely many extreme points when GG is not definably amenable. In the general, non-definably amenable case, a description of a minimal left ideal in (𝔐G,M(𝒰),)\left(\mathfrak{M}_{G,M}(\mathcal{U}),\ast\right) was obtained under some additional strong assumptions (see [CG23, Theorem 6.11] and discussion at the end of Section 4.2). Here we obtain a description of a minimal left ideal of (𝔐G,M(𝒰),)\left(\mathfrak{M}_{G,M}(\mathcal{U}),\ast\right) for an arbitrary countable NIP group:

Theorem 1.7 (Corollary 5.18).

Assume that GG is group definable in a countable NIP structure MM, and let \mathcal{M} be a minimal left ideal in (SG,M(𝒰),)(S_{G,M}(\mathcal{U}),*) and uu\in\mathcal{M} an idempotent. Then the ideal group uu\mathcal{M} carries a canonical invariant Keisler measure μu\mu_{u\mathcal{M}} (see Proposition 4.15 for the definition), and 𝔐()μu\mathfrak{M}(\mathcal{M})*\mu_{u\mathcal{M}} is a minimal left ideal of (𝔐G,M(𝒰),)\left(\mathfrak{M}_{G,M}(\mathcal{U}),\ast\right), where 𝔐()\mathfrak{M}(\mathcal{M}) denotes the space of all measures supported on \mathcal{M}, and μu\mu_{u\mathcal{M}} is an idempotent in 𝔐()μu\mathfrak{M}(\mathcal{M})*\mu_{u\mathcal{M}}.

Theorem 1.7 is deduced using a combination of two results of independent interest that we now discuss.

An important general fact from topological dynamics is that the ideal group uu\mathcal{M} of the Ellis semigroup of any flow is always a compact T1T_{1} (not necessarily Hausdorff) semi-topological group (i.e. multiplication is separately continuous) with respect to a canonical topology, the so called τ\tau-topology (which is weaker than the induced topology from the Ellis semigroup). This topology was defined by Ellis and has played an essential role in the most important structural results in abstract topological dynamics, starting from the Furstenberg structure theorem for minimal distal flows (e.g. see [Aus88]) and ending with a recent theorem of Glasner on the structure of tame, metrizable, minimal flows [Gla18]. In model theory, the τ\tau-topology on the ideal groups played a key role in applications to the quotients of definable groups by their model-theoretic connected components ([KP17]) and to Lascar strong types and quotients by arbitrary bounded invariant equivalence relations [KPR18, KR20]. It also partly motivated the work of Hrushovski on definability patterns structures with spectacular applications to additive combinatorics [Hru19, Hru20]. In [KP23], the τ\tau-topology was used to give a shorter and simpler proof of the main result of [Hru20]. As the key result of Section 4 we demonstrate the following:

Theorem 1.8.

(Lemma 4.14) Assume that GG is group definable in an arbitrary NIP structure MM, and that the τ\tau-topology on the ideal group uu\mathcal{M} of the G(M)G(M)-flow SG,M(𝒰)S_{G,M}(\mathcal{U}) is Hausdorff. Then for any clopen subset CC of SG(𝒰)S_{G}(\mathcal{U}), the subset CuC\cap u\mathcal{M} of uu\mathcal{M} is constructible, and so Borel, in the τ\tau-topology.

It follows that when the τ\tau-topology on uu\mathcal{M} is Hausdorff, the ideal group uu\mathcal{M} is a compact topological group (Corollary 4.9), so we have the unique (normalized) Haar measure huh_{u\mathcal{M}} on Borel subsets, and by Theorem 1.8 it induces the aforementioned Keisler measure μu𝔐G(𝒰)\mu_{u\mathcal{M}}\in\mathfrak{M}_{G}(\mathcal{U}) via μu(φ(x)):=hu([φ(x)]u)\mu_{u\mathcal{M}}(\varphi(x)):=h_{u\mathcal{M}}([\varphi(x)]\cap u\mathcal{M}) (Proposition 4.15). This reduces the question to understanding when the τ\tau-topology is Hausdorff — which is precisely the revised version of Newelski’s conjecture (see [KP23, Conjecture 5.3]).

The Ellis group conjecture of Newelski [New09] is an important prediction in the study of NIP groups connecting a canonical model-theoretic quotient of a definable group G(M)G(M) and a dynamical invariant of its natural action on SG,M(𝒰)S_{G,M}(\mathcal{U}). Let GG be a group definable in a structure MM, and let uu\mathcal{M} be the ideal (Ellis) group of the G(M)G(M)-flow (G(M),SG,M(𝒰))(G(M),S_{G,M}(\mathcal{U})). We let GM00G^{00}_{M} be the smallest type-definable over MM subgroup of G(𝒰)G(\mathcal{U}) of bounded index. The Ellis group conjecture of Newelski says that the group epimorphism θ:uG(𝒰)/GM00\theta:u\mathcal{M}\to G(\mathcal{U})/G^{00}_{M} given by θ(p):=a/GM00\theta(p):=a/G^{00}_{M}, for a type pup\in u\mathcal{M} and aa a realization of pp, is an isomorphism under suitable tameness assumptions on the ambient theory. This conjecture was established for definably amenable groups definable in oo-minimal structures in [CPS14], and for definably amenable groups in arbitrary NIP structures in [CS18]. On the other hand, it was refuted for SL2()\operatorname{{SL}}_{2}(\mathbb{R}) in [GPP15]. Newelski’s epimorphism θ\theta was refined in [KP17] to a sequence of epimorphisms

uu/H(u)G(𝒰)/GM000G(𝒰)/GM00,u\mathcal{M}\to u\mathcal{M}/H(u\mathcal{M})\to G(\mathcal{U})/G^{000}_{M}\to G(\mathcal{U})/G^{00}_{M},

where GM000G^{000}_{M} is the smallest bounded index subgroup of G(𝒰)G(\mathcal{U}) invariant under the action of Aut(𝒰/M)\operatorname{Aut}(\mathcal{U}/M), and H(u)H(u\mathcal{M}) is the subgroup of uu\mathcal{M} given by the intersection of the τ\tau-closures of all τ\tau-neighborhoods of uu. With this refinement, Newelski’s conjecture fails when GM000GM00G^{000}_{M}\neq G^{00}_{M}, in which case also u/H(u)G(𝒰)/GM000u\mathcal{M}/H(u\mathcal{M})\to G(\mathcal{U})/G^{000}_{M} is not an isomorphism. The first such example, the universal cover SL2()~\widetilde{\operatorname{{SL}}_{2}(\mathbb{R})} of SL2()\operatorname{{SL}}_{2}(\mathbb{R}), was found in [CP12], and further examples were given in [GK15]. On the other hand, no examples of NIP groups with non-trivial H(u)H(u\mathcal{M}) (equivalently, with uu\mathcal{M} not Hausdorff in the τ\tau-topology) were known. This motivated the following weakening of Newelski’s conjecture:

Conjecture 1.9.

[KP23, Conjecture 5.3] If MM is NIP, then the τ\tau-topology on uu\mathcal{M} is Hausdorff.

It clearly holds whenever uu\mathcal{M} is finite. It is also known to hold for definably amenable groups in NIP theories, as the full Newelski’s conjecture holds in this context. Besides those two general situations, it was confirmed only for SL2()~\widetilde{\operatorname{{SL}}_{2}(\mathbb{R})} (we refer to [KP23, Section 5] for a proof and a more detailed discussion).

In Section 5 we establish the revised Newelski’s conjecture for countable NIP groups:

Theorem 1.10 (Theorem 5.17).

The revised Newelski’s conjecture holds when GG is a definable group in a countable NIP structure.

This relies on the fundamental theorem of Glasner describing the structure of minimal tame metrizable flows [Gla18] and a presentation of the G(M)G(M)-flow SG,M(𝒰)S_{G,M}(\mathcal{U}) as the inverse limit of all SG,Δ(M)S_{G,\Delta}(M) (the Stone space of the G(M)G(M)-algebra generated by the finite set Δ\Delta), where Δ\Delta ranges over all finite collections of externally definable subsets of G(M)G(M).

2. Idempotent generically stable types

Throughout this section, we let TT be a complete theory, MTM\models T, and 𝒰M\mathcal{U}\succ M a monster model.

2.1. Generically stable types

We will need some basic facts about generically stable types. As usual, given a global type pSx(𝒰)p\in S_{x}(\mathcal{U}) (automorphism-)invariant over a small set A𝒰A\subseteq\mathcal{U} and 𝒰𝒰\mathcal{U}^{\prime}\succ\mathcal{U} a bigger monster model with respect to 𝒰\mathcal{U}, we let p|𝒰p|_{\mathcal{U}^{\prime}} be the unique extension of pp to a type in Sx(𝒰)S_{x}(\mathcal{U}^{\prime}) which is invariant over AA. Given another AA-invariant type qSy(𝒰)q\in S_{y}(\mathcal{U}), we define the AA-invariant type pqSxy(𝒰)p\otimes q\in S_{xy}(\mathcal{U}) via pq:=tp(ab/𝒰)p\otimes q:=\operatorname{tp}(ab/\mathcal{U}) for some/any a,ba,b in 𝒰\mathcal{U}^{\prime} such that bqb\models q and ap|𝒰ba\models p|_{\mathcal{U}b}. Given an arbitrary linear order (I,<)(I,<), a sequence a¯=(ai:iI)\bar{a}=(a_{i}:i\in I) in 𝒰\mathcal{U} is a Morley sequence in pp over AA if aip|Aa<ia_{i}\models p|_{Aa_{<i}} for all iIi\in I. Then the sequence a¯\bar{a} is indiscernible over AA, and for any other Morley sequence a¯=(ai:iI)\bar{a}^{\prime}=(a^{\prime}_{i}:i\in I) in pp over AA we have tp(a¯/A)=tp(a¯/A)\operatorname{tp}(\bar{a}/A)=\operatorname{tp}(\bar{a}^{\prime}/A). We can then define a global AA-invariant type p(I)((xi:iI))Sx¯(𝒰)p^{(I)}((x_{i}:i\in I))\in S_{\bar{x}}(\mathcal{U}) as

{tp(a¯/B):AB𝒰 small,a¯=(ai:iI) a Morley sequence in p over B}.\bigcup\left\{\operatorname{tp}(\bar{a}/B):A\subseteq B\subseteq\mathcal{U}\textrm{ small},\bar{a}=(a_{i}:i\in I)\textrm{ a Morley sequence in }p\textrm{ over }B\right\}.

Equivalently, p(I)=tp(a¯/𝒰)p^{(I)}=\operatorname{tp}(\bar{a}/\mathcal{U}) for a¯=(ai:iI)\bar{a}=(a_{i}:i\in I) a Morley sequence in p|𝒰p|_{\mathcal{U}^{\prime}} over 𝒰\mathcal{U}, where 𝒰𝒰\mathcal{U}^{\prime}\succ\mathcal{U} be a monster model with respect to 𝒰\mathcal{U} and p|𝒰p|_{\mathcal{U}^{\prime}} is the unique extension of pp to a type in Sx(𝒰)S_{x}(\mathcal{U}^{\prime}) which is invariant over AA. For any kωk\in\omega (viewed as an ordinal) we have p(k)(x1,,xk)=p(xk)p(x1)p^{(k)}(x_{1},\ldots,x_{k})=p(x_{k})\otimes\ldots\otimes p(x_{1}).

We do not assume NIP unless explicitly stated, and use the standard definition from [PT11]: a global type pSx(𝒰)p\in S_{x}(\mathcal{U}) is generically stable if it is AA-invariant for some small A𝒰A\subset\mathcal{U}, and for any ordinal α\alpha (or just for α=ω+ω\alpha=\omega+\omega), (ai:iα)(a_{i}:i\in\alpha) a Morley sequence in pp over AA and formula φ(x)(𝒰)\varphi(x)\in\mathcal{L}(\mathcal{U}), the set {iα:φ(ai)}\{i\in\alpha:\models\varphi(a_{i})\} is either finite or co-finite.

In the following fact, items (1)–(4) can be found in [Cas11, Section 9], (5) in [GOU13, Theorem 2.4], (6) is an immediate consequence of stationarity in (2), and (7) is [PT11, Proposition 2.1]. We let \mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}} denote forking independence. We will freely use some of the basic properties of forking in arbitrary theories, e.g. extension and left transitivity, see [CK12, Section 2] for a reference.

Fact 2.1.

Let pSx(𝒰)p\in S_{x}(\mathcal{U}) be a generically stable type, invariant over a small subset A𝒰A\subseteq\mathcal{U}. Then the following hold.

  1. (1)

    Every realization of p(ω)|Ap^{(\omega)}{|_{A}} is a totally indiscernible sequence over AA.

  2. (2)

    The type pp is the unique global non-forking extension of p|Ap|_{A}.

  3. (3)

    For any ap|Aa\models p|_{A} and bb in 𝒰\mathcal{U} such that tp(b/A)\operatorname{tp}(b/A) does not fork over AA, we have aAbbAaa\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{A}b\iff b\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{A}a (this holds for any bb when AA is an extension base, e.g. when A𝒰A\prec\mathcal{U}).

  4. (4)

    In particular, if a,bp|Aa,b\models p|_{A}, then aAb(a,b)p(2)|A(b,a)p(2)|Aa\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{A}b\iff(a,b)\models p^{(2)}|_{A}\iff(b,a)\models p^{(2)}|_{A}.

  5. (5)

    If AA is an extension base, (ai)i<ωp(ω)|A(a_{i})_{i<\omega}\models p^{(\omega)}|_{A} and φ(x,a0)\varphi(x,a_{0}) (where φ(x,y)(A)\varphi(x,y)\in\mathcal{L}(A)) forks/divides over AA, then {φ(x,ai):i<ω}\{\varphi(x,a_{i}):i<\omega\} is inconsistent.

  6. (6)

    Let ap|Aa\models p|_{A} and let b,cb,c be arbitrary small tuples in 𝒰\mathcal{U}. If aAba\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{A}b and aAbca\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{Ab}c, then aAbca\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{A}bc;

  7. (7)

    pp is definable over AA.

Remark 2.2.

By Fact 2.1(1), for a global generically stable type pp invariant over AA, we will also be using (without mentioning) an equivalent definition of a Morley sequence with the reversed order. That is, given a linear order (I,<)(I,<), we might say that (ai:iI)(a_{i}:i\in I^{*}) is a Morley sequence in pp over AA if aip|Aa<ia_{i}\models p|_{Aa_{<i}} for all iIi\in I, where II^{*} is II with the reversed ordering. So e.g. we will refer to (ak,ak1,,a1)(a_{k},a_{k-1},\ldots,a_{1}) with aip|Aa1ai1a_{i}\models p|_{Aa_{1}\ldots a_{i-1}} as a Morley sequence in pp over AA, and write (ak,,a1)p(k)(a_{k},\ldots,a_{1})\models p^{(k)}. We will frequently use this without further mention.

Fact 2.3.

[DK12, Proposition 1.2] Let pSx(𝒰)p\in S_{x}(\mathcal{U}) be a generically stable type, invariant over a small set of parameters A𝒰A\subseteq\mathcal{U}. Suppose that 𝒰𝒰\mathcal{U}\prec\mathcal{U}^{\prime}, aa is an element of 𝒰\mathcal{U}^{\prime} such that apa\models p, and bdcl(a,A)b\in\operatorname{dcl}(a,A). Then tp(b/𝒰)\operatorname{tp}(b/\mathcal{U}) is generically stable over AA.

2.2. Setting

Let G=G(x)G=G(x) be an \emptyset-type-definable group (in the sort of 𝒰\mathcal{U} corresponding to the tuple of variables xx) and G¯:=G(𝒰)\bar{G}:=G(\mathcal{U}). By \cdot we mean an \emptyset-definable function (from (𝒰x)2\left(\mathcal{U}^{x}\right)^{2} to 𝒰x\mathcal{U}^{x}) whose restriction to GG is the group operation on GG. Similarly, by -1 we mean an \emptyset-definable function (from 𝒰x\mathcal{U}^{x} to 𝒰x\mathcal{U}^{x}) whose restriction to GG is the inverse in GG. By compactness, we can fix a formula φ0(x)\varphi_{0}(x)\in\mathcal{L} implied by the partial type G(x)G(x) such that: \cdot is defined and associative on φ0(𝒰)\varphi_{0}(\mathcal{U}); ae=a=eaa\cdot e=a=e\cdot a and aa1=a1a=ea\cdot a^{-1}=a^{-1}\cdot a=e for all aφ0(𝒰)a\in\varphi_{0}(\mathcal{U}); if b1b2b_{1}\neq b_{2} then ab1ab2,b1ab2aa\cdot b_{1}\neq a\cdot b_{2},b_{1}\cdot a\neq b_{2}\cdot a for all a,b1,b2φ0(𝒰)a,b_{1},b_{2}\in\varphi_{0}(\mathcal{U}) (but φ0(𝒰)\varphi_{0}(\mathcal{U}) is not necessarily closed under \cdot). As usual, for A𝒰A\subseteq\mathcal{U}, SG(A)S_{G}(A) denotes the set of types pS(A)p\in S(A) concentrated on GG, i.e. such that p(x)G(x)p(x)\vdash G(x).

Given p,qSG(𝒰)p,q\in S_{G}(\mathcal{U}) global MM-invariant types (M𝒰M\prec\mathcal{U}), we define pqSG(𝒰)p\ast q\in S_{G}(\mathcal{U}) via pq(φ(x)):=pxqy(φ(xy))p\ast q(\varphi(x)):=p_{x}\otimes q_{y}(\varphi(x\cdot y)) for all φ(x)(U)\varphi(x)\in\mathcal{L}(U). Together with this operation, the set of all global MM-invariant types in SG(𝒰)S_{G}(\mathcal{U}) forms a left-continuous semigroup. We say that an invariant type pSG(𝒰)p\in S_{G}(\mathcal{U}) is idempotent if pp=pp\ast p=p.

2.3. Generically stable groups

Definition 2.4.

[PT11, Definition 2.1] A type-definable group G(x)G(x) is generically stable if there is a generically stable pSG(𝒰)p\in S_{G}(\mathcal{U}) which is left G(𝒰)G(\mathcal{U})-invariant (we might use “G(𝒰)G(\mathcal{U})-invariant” and “GG-invariant” interchangeably when talking about global types).

Fact 2.5.

[PT11, Lemma 2.1] Suppose that GG is a generically stable type-definable group in an arbitrary theory, witnessed by a generically stable type pSG(𝒰)p\in S_{G}(\mathcal{U}). Then we have:

  1. (1)

    pp is the unique left G(𝒰)G(\mathcal{U})-invariant and also the unique right G(𝒰)G(\mathcal{U})-invariant type;

  2. (2)

    p=p1p=p^{-1} (where p1:=tp(g1/𝒰)p^{-1}:=\operatorname{tp}(g^{-1}/\mathcal{U}) for some/any gpg\models p in a bigger monster model 𝒰𝒰\mathcal{U}^{\prime}\succ\mathcal{U}).

By Fact 2.4 and its symmetric version, we get:

Corollary 2.6.

A type-definable group G(x)G(x) is generically stable if and only if there is a generically stable pSG(𝒰)p\in S_{G}(\mathcal{U}) which is right G(𝒰)G(\mathcal{U})-invariant.

2.4. Idempotent generically stable types and generic transitivity: main conjecture

Let pSG(𝒰)p\in S_{G}(\mathcal{U}) be a generically stable type over MM. The following is standard:

Proposition 2.7.

The left stabilizer Stab(p)\operatorname{Stab}(p) of pp is an intersection of relatively MM-definable subgroups of G¯\bar{G}; in particular, it is MM-type-definable.

Proof.

By compactness, we can find a formula φ1(x)\varphi_{1}(x)\in\mathcal{L} implied by G(x)G(x) such that φ1(G¯)G¯φ0(G¯)\varphi_{1}(\bar{G})\cdot\bar{G}\subseteq\varphi_{0}(\bar{G}).

Since pp is generically stable over MM, it is definable over MM (by Fact 2.1(7)). For any formula φ(x,y)\varphi(x,y)\in\mathcal{L} which implies φ1(x)\varphi_{1}(x) (and yy is an arbitrary tuple of variables) let

dpφ:={(h,a)φ0(𝒰)×𝒰y:φ(hx,a)p}.d_{p}\varphi:=\left\{(h,a)\in\varphi_{0}(\mathcal{U})\times\mathcal{U}^{y}:\varphi(h\cdot x,a)\in p\right\}.

By the definability of pp over MM, dpφd_{p}\varphi is definable over MM. Then

Stabφ(p):={gG¯:gdpφ=dpφ},\operatorname{Stab}_{\varphi}(p):=\{g\in\bar{G}:g\cdot d_{p}\varphi=d_{p}\varphi\},

where g(h,a):=(hg1,a)g\cdot(h,a):=(h\cdot g^{-1},a), is an MM-relatively definable subgroup of G¯\bar{G} (the fact that it is a subgroup of GG follows from the observation that it is the stabilizer of the set dpφd_{p}\varphi under the left action of GG on φ0(𝒰)G¯×𝒰y\varphi_{0}(\mathcal{U})\cdot\bar{G}\times\mathcal{U}^{y} given by g(h,a):=(hg1,a)g\cdot(h,a):=(h\cdot g^{-1},\ a), which uses the choice of φ0(x)\varphi_{0}(x)). By the choice of φ1(x)\varphi_{1}(x), we get that Stab(p)=φ(x,y),φ(x,y)φ1(x)Stabφ(p)\operatorname{Stab}(p)=\bigcap_{\varphi(x,y)\in\mathcal{L},\varphi(x,y)\vdash\varphi_{1}(x)}\operatorname{Stab}_{\varphi}(p) is an intersection of relatively MM-definable subgroups of G¯\bar{G}. ∎

Example 2.8.

Let GG^{\prime} be an arbitrary type-definable subgroup of GG which is generically stable, witnessed by a generically stable left or right GG^{\prime}-invariant type pSG(𝒰)p\in S_{G^{\prime}}(\mathcal{U}). Then pp is obviously idempotent.

Our central question in the case of types is whether this is the only source of generically stable idempotent types:

Definition 2.9.

For the rest of the section, we let H:=Stab(p)H_{\ell}:=\operatorname{Stab}_{\ell}(p) and Hr:=Stabr(p)H_{\textrm{r}}:=\operatorname{Stab}_{\textrm{r}}(p) be the left and the right stabilizer of pp, respectively.

Problem 2.10.

Assume that pp is generically stable and idempotent, and let H=HH=H_{\ell} or H=HrH=H_{\textrm{r}}. Is it true that pSH(𝒰)p\in S_{H}(\mathcal{U}) and the group HH is generically stable (Definition 2.4)?

We will note now that the second part follows from the first, and that the left and the right versions of the problem are equivalent.

Remark 2.11.

Assume pp is generically stable and pSH(𝒰)p\in S_{H}(\mathcal{U}), where H=HH=H_{\ell} or H=HrH=H_{\textrm{r}}. Then:

  1. (1)

    HH is a generically stable group, witnessed by pp (hence pp is both the unique left-invariant and the unique right-invariant type of HH, by Fact 2.5);

  2. (2)

    HH is the smallest among all type-definable subgroups HH^{\prime} of GG with pSH(𝒰)p\in S_{H^{\prime}}(\mathcal{U});

  3. (3)

    HH is both the left and the right stabilizer of pp in GG.

Proof.

We will do the case of H=HH=H_{\ell}. The proof for H=HrH=H_{\textrm{r}} is symmetric.

(1) As then pp is a left HH-invariant generically stable type in SH(𝒰)S_{H}(\mathcal{U}).

(2) HH is type-definable by Proposition 2.7. For any type-definable HG(𝒰)H^{\prime}\leq G(\mathcal{U}) with p(x)H(x)p(x)\vdash H^{\prime}(x), the group H′′:=HHH^{\prime\prime}:=H\cap H^{\prime} is type-definable with p(x)H′′(x)p(x)\vdash H^{\prime\prime}(x) and H′′HH^{\prime\prime}\leq H. If the index is 2\geq 2, we have some gHg\in H with H′′gH′′=H^{\prime\prime}\cap g\cdot H^{\prime\prime}=\emptyset, and p(x)H′′(x),(gp)(x)(gH′′)(x)p(x)\vdash H^{\prime\prime}(x),\left(g\cdot p\right)(x)\vdash\left(g\cdot H^{\prime\prime}\right)(x), so pgpp\neq g\cdot p — a contradiction. So H′′=HH^{\prime\prime}=H, and HHH\subseteq H^{\prime}.

(3) HrH_{r} is type-definable by a symmetric argument as in Proposition 2.7. By (1), pp is right HH-invariant, so we have HHrH\subseteq H_{r}, and so pSHr(𝒰)p\in S_{H_{r}}(\mathcal{U}). By the right version of (2) (which is obtained by a symmetric argument as in (2)), we conclude that H=HrH=H_{r}. ∎

By Remark 2.11, we see that pSH(𝒰)p\in S_{H_{\ell}}(\mathcal{U}) if and only if pSHr(𝒰)p\in S_{H_{\textrm{r}}}(\mathcal{U}), so the left and the right versions of Problem 2.10 are indeed equivalent.

Let pSG(𝒰)p\in S_{G}(\mathcal{U}) be a generically stable type. Let 𝒰𝒰\mathcal{U}^{\prime}\succ\mathcal{U} be a monster model with respect to 𝒰\mathcal{U} and p:=p|𝒰p^{\prime}:=p|_{\mathcal{U}^{\prime}} (the unique extension of pp to a type in SG(𝒰)S_{G}(\mathcal{U}^{\prime}) which is invariant over MM). Then still pp^{\prime} is definable over MM and H(𝒰)=Stab(p)H(\mathcal{U}^{\prime})=\operatorname{Stab}(p^{\prime}). Pick an arbitrary ap(𝒰)a\in p(\mathcal{U}^{\prime}).

Let pSG(𝒰)p\in S_{G}(\mathcal{U}) be a generically stable type, and let 𝒰𝒰\mathcal{U}^{\prime}\succ\mathcal{U} be a monster model with respect to 𝒰\mathcal{U} and p:=p|𝒰p^{\prime}:=p|_{\mathcal{U}^{\prime}} (the unique extension of pp to a type in SG(𝒰)S_{G}(\mathcal{U}^{\prime}) which is invariant over MM). Then still pp^{\prime} is definable over MM and H(𝒰)=Stab(p)H_{\ell}(\mathcal{U}^{\prime})=\operatorname{Stab}_{\ell}(p^{\prime}), Hr(𝒰)=Stabr(p)H_{\textrm{r}}(\mathcal{U}^{\prime})=\operatorname{Stab}_{\textrm{r}}(p^{\prime}). Pick an arbitrary ap(𝒰)a\in p(\mathcal{U}^{\prime}).

Remark 2.12.

The following conditions are equivalent for a generically stable type pp:

  1. (1)

    pSH(𝒰)p\in S_{H_{\ell}}(\mathcal{U});

  2. (2)

    aStab(p)a\in\operatorname{Stab}_{\ell}(p^{\prime});

  3. (3)

    for any (equivalently, some) (a0,a1)p(2)(a_{0},a_{1})\models p^{(2)}, (a0a1,a0)p(2)(a_{0}\cdot a_{1},a_{0})\models p^{(2)};

  4. (4)

    pSHr(𝒰)p\in S_{H_{\textrm{r}}}(\mathcal{U});

  5. (5)

    aStabr(p)a\in\operatorname{Stab}_{\textrm{r}}(p^{\prime});

  6. (6)

    for any (equivalently, some) (a0,a1)p(2)(a_{0},a_{1})\models p^{(2)}, (a1a0,a0)p(2)(a_{1}\cdot a_{0},a_{0})\models p^{(2)}.

Proof.

(1)(2)(1)\Rightarrow(2). This is because ap(𝒰)H(𝒰)=Stab(p)a\in p(\mathcal{U}^{\prime})\subseteq H_{\ell}(\mathcal{U}^{\prime})=\operatorname{Stab}_{\ell}(p^{\prime}), where the inclusion follows by (1).

(2)(3)(2)\Rightarrow(3). Pick bpb\models p^{\prime}. Then (a,b)p(2)(a,b)\models p^{(2)} and abpa\cdot b\models p^{\prime} by (2). So (ab,a)p(2)(a\cdot b,a)\models p^{(2)}.

(3)(1)(3)\Rightarrow(1). Suppose (1) fails. Then aH(𝒰)=Stab(p)a\notin H_{\ell}(\mathcal{U}^{\prime})=\operatorname{Stab}_{\ell}(p^{\prime}). So for any bpb\models p^{\prime} we have that aba\cdot b does not realize pp^{\prime}. Using Fact 2.1, this implies that aba\cdot b does not realize p|𝒰ap|_{\mathcal{U}a}, because otherwise, since ab|𝒰a𝒰a\cdot b\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{\mathcal{U}a}\mathcal{U}^{\prime} (which follows from b|𝒰a𝒰b\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{\mathcal{U}a}\mathcal{U}^{\prime} and left transitivity of forking), tp(ab/𝒰)\operatorname{tp}(a\cdot b/\mathcal{U}^{\prime}) is the unique non-forking extension of p|𝒰ap|_{\mathcal{U}a} which is exactly pp^{\prime}, a contradiction. Thus, (ab,a)(a\cdot b,a) does not realize p(2)p^{(2)} which contradicts (3).

(4)(5)(6)(4)(4)\Rightarrow(5)\Rightarrow(6)\Rightarrow(4) are obtained by symmetric arguments to the above ones.

(1)(4)(1)\Leftrightarrow(4) follows from Remark 2.11. ∎

Definition 2.13.

We will say that a generically stable type pSG(𝒰)p\in S_{G}(\mathcal{U}) is generically transitive if it satisfies any of the equivalent conditions in Remark 2.12.

Remark 2.14.

We have chosen this terminology to highlight the connection of condition (3) in Remark 2.12(3) and the generic transitivity assumption in Hrushovski’s group chunk theorem (see e.g. [Bay18, Section 5.1]).

In view of Remark 2.12, our main Problem 2.10 is equivalent to the following:

Problem 2.15.

Assume that pp is generically stable and idempotent. Is it then generically transitive?

In the following sections, we will provide a positive solution under some additional assumptions on the group. We will also see that elements of stable group theory (stratified rank, connected components, etc.) can be developed in an arbitrary theory localizing on a generically stable type pp, if and only said type pp is generically transitive.

We conclude this section with a couple of additional observations.

Lemma 2.16.

If pSG(𝒰)p\in S_{G}(\mathcal{U}) is generically stable over MM and idempotent, then the type p1:=tp(a1/𝒰)p^{-1}:=\operatorname{tp}(a^{-1}/\mathcal{U}) for some/any apa\models p is also generically stable over MM and idempotent.

Proof.

Assume the hypothesis. Then p1p^{-1} is generically stable over MM by Fact 2.3. It follows that p1p1p^{-1}*p^{-1} is definable over MM (see e.g. [CG22, Proposition 3.15]). Since p1p^{-1} is the unique extension of p1|Mp^{-1}|_{M} definable over MM, it suffices to show that (p1p1)|M=p1|M(p^{-1}*p^{-1})|_{M}=p^{-1}|_{M}. Let b1p1|Mb_{1}\models p^{-1}|_{M} and b2p1|Mb1b_{2}\models p^{-1}|_{Mb_{1}}. Clearly b11p|Mb_{1}^{-1}\models p|_{M} and b21p|Mb1b_{2}^{-1}\models p|_{Mb_{1}} and so (b21,b11)p(2)|M(b_{2}^{-1},b_{1}^{-1})\models p^{(2)}|_{M}. Since pp is idempotent, we conclude that b11b21p|Mb_{1}^{-1}\cdot b_{2}^{-1}\models p|_{M} and in particular (b2b1)p1|M(b_{2}\cdot b_{1})\models p^{-1}|_{M}. For any φ(x)x(M)\varphi(x)\in\mathcal{L}_{x}(M) we have

φ(x)p1p1φ(xy)px1py1φ(xb1)px1\varphi(x)\in p^{-1}*p^{-1}\implies\varphi(x\cdot y)\in p^{-1}_{x}\otimes p^{-1}_{y}\implies\varphi(x\cdot b_{1})\in p^{-1}_{x}
φ(b2b1)φ(x)tp(b2b1/M)φ(x)p1|M,\implies\models\varphi(b_{2}\cdot b_{1})\implies\varphi(x)\in\operatorname{tp}(b_{2}\cdot b_{1}/M)\implies\varphi(x)\in p^{-1}|_{M},

hence (p1p1)|M=p1|M(p^{-1}*p^{-1})|_{M}=p^{-1}|_{M}. ∎

Remark 2.17.

If pp is generically transitive, then p=p1p=p^{-1}.

Indeed, by Remarks 2.12 and 2.11 it follows that if pp is generically transitive, then pSH(𝒰)p\in S_{H_{\ell}}(\mathcal{U}) witnesses that HH_{\ell} is generically stable, hence p=p1p=p^{-1} by Fact 2.5.

2.5. Abelian groups

In this section we give a positive answer to Problem 2.10 in the case of abelian groups (in arbitrary theories):

Proposition 2.18.

Assume that GG is an abelian group and pSG(𝒰)p\in S_{G}(\mathcal{U}) is a generically stable idempotent type. Then pp is generically transitive.

Proof.

Let M𝒰M\prec\mathcal{U} be any small model such that pp is MM-invariant. Let (a1,a0)p(2)|M(a_{1},a_{0})\models p^{(2)}|_{M} be given, and assume towards contradiction that (a1a0,a0)⊧̸p(2)|M(a_{1}\cdot a_{0},a_{0})\not\models p^{(2)}|_{M}, then a1a0/Ma0a_{1}\cdot a_{0}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{M}a_{0} by Fact 2.1(4). Let φ(x,y)(M)\varphi(x,y)\in\mathcal{L}(M) be such that φ(a1a0,a0)\models\varphi(a_{1}\cdot a_{0},a_{0}) and φ(x,a0)\varphi(x,a_{0}) forks over MM. We extend (a1,a0)(a_{1},a_{0}) to a Morley sequence (ai)i<ωp(ω)|M(a_{i})_{i<\omega}\models p^{(\omega)}|_{M}. For each k<ωk<\omega, let bk:=ak1ak2a0b_{k}:=a_{k-1}\cdot a_{k-2}\cdot\ldots\cdot a_{0}. Then we have:

  1. (1)

    (ak1,,a1)p(k1)|a0M(a_{k-1},\ldots,a_{1})\models p^{(k-1)}|_{a_{0}M}, hence by idempotence of pp we have ak1a1p|a0Ma_{k-1}\cdot\ldots\cdot a_{1}\models p|_{a_{0}M}, and so (ak1a1,a0)p(2)|M(a_{k-1}\cdot\ldots\cdot a_{1},a_{0})\models p^{(2)}|_{M}, and bk=(ak1a1)a0b_{k}=(a_{k-1}\cdot\ldots\cdot a_{1})\cdot a_{0}, so (bk,a0)M(a1a0,a0)(b_{k},a_{0})\equiv_{M}(a_{1}\cdot a_{0},a_{0}), in particular φ(bk,a0)\models\varphi(b_{k},a_{0});

  2. (2)

    for any permutation σ\sigma of {0,1,,k1}\{0,1,\ldots,k-1\}, we have (aσ(0),,aσ(k1))p(k)|M(a_{\sigma(0)},\ldots,a_{\sigma(k-1)})\models p^{(k)}|_{M} (by Fact 2.1(1));

  3. (3)

    in particular, for every i<ki<k we have

    (ak1,ak2,,a0)M(ak1,,ai+1,ai1,,a0,ai);(a_{k-1},a_{k-2},\ldots,a_{0})\equiv_{M}(a_{k-1},\ldots,a_{i+1},a_{i-1},\ldots,a_{0},a_{i});
  4. (4)

    and bk=ak1ai+1ai1a0aib_{k}=a_{k-1}\cdot\ldots\cdot a_{i+1}\cdot a_{i-1}\cdot\ldots\cdot a_{0}\cdot a_{i} (as 𝒢\mathcal{G} is abelian);

  5. (5)

    hence (bk,a0)M(bk,ai)(b_{k},a_{0})\equiv_{M}(b_{k},a_{i}) for every i<ki<k.

Thus, by (1) and (5), for every k<ωk<\omega we have bk{φ(x,ai):i<k}b_{k}\models\{\varphi(x,a_{i}):i<k\}, and by compactness the set {φ(x,ai):i<ω}\{\varphi(x,a_{i}):i<\omega\} is consistent. But this contradicts the choice of φ\varphi by Fact 2.1(5). Hence (a1a0,a0)p(2)|M(a_{1}\cdot a_{0},a_{0})\models p^{(2)}|_{M}, and since MM was an arbitrary small model (over which pp is invariant), we conclude the proof. ∎

Remark 2.19.

It was pointed out to us by Martin Hils that this argument is related to [HRK19, Lemma 5.1], which is used there to find idempotent types in abelian groups of finite pp-weight.

Remark 2.20.

In the case of an arbitrary group, the proof of Proposition 2.18 gives the following:

  • If pSG(𝒰)p\in S_{G}(\mathcal{U}) is invariant and idempotent, then for any (a2,a1,a0)p(3)(a_{2},a_{1},a_{0})\models p^{(3)} we have (a2a1a0,a1)p(2)(a_{2}\cdot a_{1}\cdot a_{0},a_{1})\models p^{(2)}.

Indeed, this time we assume towards a contradiction that a2a1a0/Ma1a_{2}\cdot a_{1}\cdot a_{0}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{M}a_{1}. We extend (a2,a1,a0)(a_{2},a_{1},a_{0}) to a Morley sequence (ai:iω)(a_{i}:i\in\omega) in pp over MM. Let kωk\in\omega be arbitrary, and let bk:=aka0b_{k}:=a_{k}\cdot\ldots\cdot a_{0}. Then we get (bk,ai)M(a2a1a0,a1)(b_{k},a_{i})\equiv_{M}(a_{2}\cdot a_{1}\cdot a_{0},a_{1}) for all 2ik12\leq i\leq k-1, which gives a contradiction as in the proof of Proposition 2.18. To see this, note that since pp is idempotent, a0:=ai1a0p|Ma_{0}^{\prime}:=a_{i-1}\cdot\ldots\cdot a_{0}\models p|_{M}, aip|Ma0a_{i}\models p|_{Ma_{0}^{\prime}}, and a2:=akai+1p|Ma0aia_{2}^{\prime}:=a_{k}\cdot\ldots\cdot a_{i+1}\models p|_{Ma_{0}^{\prime}a_{i}}. Hence (a2,ai,a0)p(3)|M(a_{2}^{\prime},a_{i},a_{0}^{\prime})\models p^{(3)}|_{M}, and so (bk,ai)M(a2a1a0,a1)(b_{k},a_{i})\equiv_{M}(a_{2}\cdot a_{1}\cdot a_{0},a_{1}).

Remark 2.21.

Note that in Remark 2.20 we only assumed that the idempotent type pp is invariant. However, the assumption of generic stability is necessary in Proposition 2.18. Indeed, let M:=(,+,<)M:=(\mathbb{R},+,<), G(M):=(,+)G(M):=(\mathbb{R},+), and let p0+SG(𝒰)p_{0^{+}}\in S_{G}(\mathcal{U}) be the unique global definable (over \mathbb{R}) type extending {x<a:a𝒰,a>0}{x>a:a𝒰,a0}\{x<a:a\in\mathcal{U},a>0\}\cup\{x>a:a\in\mathcal{U},a\leq 0\}. Then p0+p_{0^{+}} is idempotent (see [CG23, Example 4.5(1)]). But if (a1,a0)p0+p0+(a_{1},a_{0})\models p_{0^{+}}\otimes p_{0^{+}} we have 0<a1<a0<𝒰0<a_{1}<a_{0}<\mathcal{U}, hence a1+a0>a0a_{1}+a_{0}>a_{0}, so (a1+a0,a0)⊧̸p0+p0+(a_{1}+a_{0},a_{0})\not\models p_{0^{+}}\otimes p_{0^{+}}.

2.6. Inp-minimal groups

A similar argument with (local) weight applies to arbitrary inp-minimal groups. Recall that a type-definable group GG is inp-minimal if bdn(G(x))1\operatorname{bdn}(G(x))\leq 1 (we refer to [Adl07] and [Che14, Section 2] for the definition and basic properties of burden). In particular, every dp-minimal group is inp-minimal. Note that there exist dp-minimal groups which are not virtually abelian [Sim03]. Answering [CPS14, Problem 5.9], it was recently proved in [Sto23] and [Wag24] that all dp-minimal groups are virtually nilpotent.

We will use the following fact:

Fact 2.22.

[GOU13, Theorem 2.9] Let p(x)S(M)p(x)\in S(M) be generically stable over a small set AA and (ai:i<κ)p(κ)|A(a_{i}:i<\kappa)\models p^{(\kappa)}|_{A} for some cardinal κ\kappa. Let b𝒰yb\in\mathcal{U}^{y} be such that b/Aaib\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mathchar 12854\relax$\kern 8.00134pt\hss}\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{A}a_{i} for all i<κi<\kappa. Then bdn(tp(b/A))κ\operatorname{bdn}(\operatorname{tp}(b/A))\geq\kappa.

Proposition 2.23.

Assume that GG is an arbitrary type-definable group which is inp-minimal. If pSG(𝒰)p\in S_{G}(\mathcal{U}) is idempotent and generically stable, then pp is generically transitive.

Proof.

Assume that GG is inp-minimal and pSG(𝒰)p\in S_{G}(\mathcal{U}) is a generically stable idempotent type. Let 𝒰𝒰\mathcal{U}^{\prime}\succ\mathcal{U} be a monster model with respect to 𝒰\mathcal{U} and p:=p|𝒰p^{\prime}:=p|_{\mathcal{U}^{\prime}} (the unique extension of pp to a type in SG(𝒰)S_{G}(\mathcal{U}^{\prime}) which is invariant over MM). Then pSG(𝒰)p^{\prime}\in S_{G}(\mathcal{U}^{\prime}) is idempotent and generically stable over 𝒰\mathcal{U}.

Let (a1,a0)p(2)=(p)(2)|𝒰(a_{1},a_{0})\models p^{(2)}=(p^{\prime})^{(2)}|_{\mathcal{U}}. Let b:=a1a0G(𝒰)b:=a_{1}\cdot a_{0}\in G(\mathcal{U}^{\prime}), by assumption bdn(tp(b/𝒰))bdn(G(x))1\operatorname{bdn}\left(\operatorname{tp}(b/\mathcal{U})\right)\leq\operatorname{bdn}(G(x))\leq 1. As a1𝒰a0a_{1}\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{\mathcal{U}}a_{0}, by Fact 2.22 we must have at least one of the following (by idempotence and Fact 2.1(4)):

  1. (1)

    b𝒰a0b\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{\mathcal{U}}a_{0}, hence (a1a0,a0)p(2)(a_{1}\cdot a_{0},a_{0})\models p^{(2)};

  2. (2)

    or b𝒰a1b\mathop{\mathchoice{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\displaystyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\textstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptstyle{}}{\kern 5.71527pt\hbox to0.0pt{\hss$\mid$\hss}\lower 3.87495pt\hbox to0.0pt{\hss$\smile$\hss}\kern 5.71527pt\scriptscriptstyle{}}}_{\mathcal{U}}a_{1}, hence (a1a0,a1)p(2)(a_{1}\cdot a_{0},a_{1})\models p^{(2)}.

In both cases we obtain that pp is generically transitive (see Remark 2.12). ∎

Remark 2.24.

Fact 2.22 also holds when pp a generically simple type in an NTP2 theory and AA is an extension base (this follows from [Che14, Section 6] and [Sim20, Section 3.1]).

2.7. Stable theories

In this section we provide a proof that all idempotent types in stable groups are generically transitive. This was known from [New91] (see also e.g. [BMPW16, Lemme 1.2] and references there), and recently generalized from idempotent types to idempotent Keisler measures in stable theories in [CG22] (see the discussion in Section 3.10). We provide two detailed proofs since in the following sections we will extend them to the case when pp is a stable type in an arbitrary theory, and also to the case of a generically stable type pp in a simple or even rosy theory.

The proof uses local stratified ranks:

Definition 2.25.

Let G(x)G(x) be an \emptyset-type-definable group and φ0(x)\varphi_{0}(x)\in\mathcal{L} as in Section 2.2. To a formula φ(x,y)\varphi(x,y)\in\mathcal{L} (with yy an arbitrary tuple of variables) we associate a formula φ(x,y):=φ(x,y)φ0(x)\varphi^{\prime}(x,y):=\varphi(x,y)\wedge\varphi_{0}(x). For gG¯g\in\bar{G}, put φg(x,y):=gφ(x,y):=(z)(φ(z,y)x=gz)\varphi_{g}(x,y):=g\cdot\varphi^{\prime}(x,y):=(\exists z)(\varphi^{\prime}(z,y)\wedge x=g\cdot z). Finally, let Δφ:={φg(x,y):gG¯}\Delta_{\varphi}:=\{\varphi_{g}(x,y):g\in\bar{G}\}. We consider the usual notion of Δφ\Delta_{\varphi}-rank denoted by RΔφR_{\Delta_{\varphi}} and Δφ\Delta_{\varphi}-multiplicity denoted by MltΔφ\operatorname{{Mlt}}_{\Delta_{\varphi}}.

The proofs of all items except (3) in the following fact are standard arguments as for usual Δ\Delta-ranks (see e.g. [Pil96, Chapter 1]). The proof of (3) uses the choice of φ0(x)\varphi_{0}(x) and is left as an exercise.

Fact 2.26.

Assume TT is stable. Then we have:

  1. (1)

    RΔφ(x=x)<ωR_{\Delta_{\varphi}}(x=x)<\omega;

  2. (2)

    RΔφ(ψ1(x)ψ2(x))=max(RΔφ(ψ1),RΔφ(ψ2))R_{\Delta_{\varphi}}(\psi_{1}(x)\lor\psi_{2}(x))=\max(R_{\Delta_{\varphi}}(\psi_{1}),R_{\Delta_{\varphi}}(\psi_{2}));

  3. (3)

    RΔφR_{\Delta_{\varphi}} is invariant under left translations by the elements of G¯\bar{G};

  4. (4)

    For any AB𝒰A\subseteq B\subset\mathcal{U} and qSG(B)q\in S_{G}(B) we have that qq does not fork over AA if and only if RΔφ(q)=RΔφ(q|A)R_{\Delta_{\varphi}}(q)=R_{\Delta_{\varphi}}(q|_{A}) for every φ\varphi\in\mathcal{L};

  5. (5)

    MltΔφ(q)=1\operatorname{{Mlt}}_{\Delta_{\varphi}}(q)=1 for any complete type qSG(N)q\in S_{G}(N) over a model N𝒰N\prec\mathcal{U}.

Remark 2.27.

Note that (5) follows from (4) and stationary of types over models in stable theories, but for 0\aleph_{0}-saturated NN it can also be shown easily directly, without using forking.

In the proof below, the role of 𝒰\mathcal{U} is played by 𝒰\mathcal{U}^{\prime}.

Proposition 2.28.

If TT is stable and pSG(𝒰)p\in S_{G}(\mathcal{U}) is idempotent, then pp is generically transitive.

Proof.

By stability, pp is generically stable over some small M𝒰M\prec\mathcal{U}. Let apa\models p in 𝒰\mathcal{U}^{\prime}. By Remark 2.12, it suffices to show that aStab(p)a\in\operatorname{Stab}_{\ell}(p^{\prime}), where p=p|𝒰p^{\prime}=p|_{\mathcal{U}^{\prime}}

Method 1 (without using forking). Suppose for a contradiction that appa\cdot p^{\prime}\neq p^{\prime}, witnessed by a formula φ(x,b)(𝒰)\varphi(x,b)\in\mathcal{L}(\mathcal{U}^{\prime}). Since pp is a generically stable idempotent and ap(𝒰)a\in p(\mathcal{U}^{\prime}), we get ap|𝒰=pa\cdot p^{\prime}|_{\mathcal{U}}=p. On the other hand, since pp^{\prime} is invariant over MM and the ranks are invariant under automorphisms, a formula ψ(x)p\psi(x)\in p^{\prime} with RΔφ(x,y)(ψ)=RΔφ(x,y)(p)R_{\Delta_{\varphi(x,y)}}(\psi)=R_{\Delta_{\varphi(x,y)}}(p^{\prime}) and MltΔφ(x,y)(ψ)=1\operatorname{{Mlt}}_{\Delta_{\varphi(x,y)}}(\psi)=1 can be mapped by an automorphism over MM to a formula ψ(x)p|𝒰=p\psi^{\prime}(x)\in p^{\prime}|_{\mathcal{U}}=p with RΔφ(x,y)(ψ)=RΔφ(x,y)(ψ)R_{\Delta_{\varphi(x,y)}}(\psi^{\prime})=R_{\Delta_{\varphi(x,y)}}(\psi) and MltΔφ(x,y)(ψ)=1\operatorname{{Mlt}}_{\Delta_{\varphi(x,y)}}(\psi^{\prime})=1. By Fact 2.26(3), we also have RΔφ(x,y)(ap)=RΔφ(x,y)(p)R_{\Delta_{\varphi(x,y)}}(a\cdot p^{\prime})=R_{\Delta_{\varphi(x,y)}}(p^{\prime}). Summarizing, ppapp\subseteq p^{\prime}\cap a\cdot p^{\prime}, RΔφ(x,y)(p)=RΔφ(x,y)(p)=RΔφ(x,y)(ap)R_{\Delta_{\varphi(x,y)}}(p)=R_{\Delta_{\varphi(x,y)}}(p^{\prime})=R_{\Delta_{\varphi(x,y)}}(a\cdot p^{\prime}) and MltΔφ(x,y)(p)=1\operatorname{{Mlt}}_{\Delta_{\varphi(x,y)}}(p)=1. Hence, p|φ(x,y)=ap|φ(x,y)p^{\prime}|_{\varphi(x,y)}=a\cdot p^{\prime}|_{\varphi(x,y)}, a contradiction.

Method 2 (using forking). Since pp^{\prime} does not fork over 𝒰\mathcal{U}, by Fact 2.26(4), we get RΔφ(p)=RΔφ(p)R_{\Delta_{\varphi}}(p^{\prime})=R_{\Delta_{\varphi}}(p) for every φ(x,y)\varphi(x,y)\in\mathcal{L}. By Fact 2.26(3), we also have RΔφ(ap)=RΔφ(p)R_{\Delta_{\varphi}}(a\cdot p^{\prime})=R_{\Delta_{\varphi}}(p^{\prime}). Thus, RΔφ(ap)=RΔφ(p)R_{\Delta_{\varphi}}(a\cdot p^{\prime})=R_{\Delta_{\varphi}}(p). On the other hand, since pp is a generically stable idempotent and ap(𝒰)a\in p(\mathcal{U}^{\prime}), we have papp\subseteq a\cdot p^{\prime}. We conclude, using Fact 2.26(4), that apa\cdot p^{\prime} does not fork over 𝒰\mathcal{U}, and so it is the unique non-forking extension of pp which is equal to pp^{\prime}. ∎

2.8. Stable types in arbitrary theories

Method 1 from the proof of Proposition 2.28 extends to the case when pSG(𝒰)p\in S_{G}(\mathcal{U}) is stable over MM, that is pp is MM-invariant and p|Mp|_{M} is stable (in a not necessarily stable theory). Recall that p(x)|Mp(x)|_{M} is stable if there are no sequences (ai)i<ω(a_{i})_{i<\omega} in 𝒰x\mathcal{U}^{x} and (bi)i<ω(b_{i})_{i<\omega} in 𝒰y\mathcal{U}^{y} such that aip|Ma_{i}\models p|_{M} for all i<ωi<\omega and for some φ(x,y)\varphi(x,y)\in\mathcal{L} we have φ(ai,bj)i<j\models\varphi(a_{i},b_{j})\iff i<j for all i,j<ωi,j<\omega. It follows from the definitions that if pp is stable over MM, then it is generically stable over MM.

We refer to e.g. [ACP14, Section 1] and [HO10, Section 2] for the basic properties of stable types. Method 1 applies when pp is stable over MM because in this case RΔφ(p)<ωR_{\Delta_{\varphi}}(p)<\omega, and we also have item (3) of Fact 2.26 (without any assumptions on TT). Thus, Method 1 yields:

Proposition 2.29.

If pSG(𝒰)p\in S_{G}(\mathcal{U}) is idempotent and stable over some small M𝒰M\prec\mathcal{U}, then pp is generically transitive.

In order to prove this proposition using Method 2 from the proof of Proposition 2.28, we have to be careful with item (4) of Fact 2.26. Modifying a standard proof of item (4) (see e.g.  [Pil96, Lemma 3.4]) gives the following weaker variant:

Fact 2.30.

Assume pSG(𝒰)p\in S_{G}(\mathcal{U}) and p|Mp|_{M} is stable for a small model M𝒰M\prec\mathcal{U}. Let BB be a set of realizations of p|Mp|_{M} and ABA\subseteq B. Then for any qSG(MB)q\in S_{G}(MB) extending p|Mp|_{M} we have that qq does not fork over MAMA if and only if RΔφ(q)=RΔφ(q|MA)R_{\Delta_{\varphi}}(q)=R_{\Delta_{\varphi}}(q|_{MA}) for every φ\varphi\in\mathcal{L}.

While adapting the proof of [Pil96, Lemma 3.4] for Fact 2.30, the only essential difficulty is to show that if a formula φ(x,B)\varphi(x,B) does not fork over MAMA (where φ(x,y)(MA)\varphi(x,y)\in\mathcal{L}(MA)), then some positive Boolean combination of MAMA-conjugates of φ(x,B)\varphi(x,B) is definable over MAMA. For that one needs to use the assumption that BB is a set of realizations of p|Mp|_{M} in order to have that tp(B/MA)\operatorname{tp}(B/MA) is stable, which allows to use symmetry of forking and acleq(MA)\operatorname{acl}^{\textit{eq}}(MA)-definability of non-forking extensions of tp(B/MA)\operatorname{tp}(B/MA).

The following is a strengthening of the fact saying that pp is the unique non-forking extension of p|Mp|_{M} to a global type, and follows by one of the standard proofs:

Proposition 2.31.

Assume that pSG(𝒰)p\in S_{G}(\mathcal{U}) is generically stable, and p=p|𝒰p^{\prime}=p|_{\mathcal{U}^{\prime}} (where 𝒰𝒰\mathcal{U}^{\prime}\succ\mathcal{U} is a monster model with respect to 𝒰\mathcal{U}). Let qSG(𝒰)q\in S_{G}(\mathcal{U}^{\prime}) be an extension of pp such that q|𝒰Bq|_{\mathcal{U}B} does not fork over 𝒰\mathcal{U} for every small set BB of independent realizations of pp. Then q=pq=p^{\prime}.

More generally, in the assumption, it is enough to consider only the sets BB of independent realizations of pp containing a fixed realization aa of pp in 𝒰\mathcal{U}^{\prime}.

Proof.

We will deduce the proposition from the following claim.

Claim.

If aiq|𝒰,a<ia_{i}\models q|_{\mathcal{U},a_{<i}} for all i<ni<n, then (a0,,an1)(a_{0},\dots,a_{n-1}) is a Morley sequence in pp. In the more general version of the assumption, the same holds but assuming that a0=aa_{0}=a.

Proof.

This is induction on nn. The base step n=0n=0 is trivial, as q|𝒰=pq|_{\mathcal{U}}=p by assumption.

Induction step. Consider any aiq|𝒰,a<ia_{i}\models q|_{\mathcal{U},a_{<i}} in 𝒰\mathcal{U}^{\prime} for all ini\leq n. By induction hypothesis aip|𝒰,a<ia_{i}\models p|_{\mathcal{U},a_{<i}} for all i<ni<n. The goal is to prove that anp|𝒰,a<na_{n}\models p|_{\mathcal{U},a_{<n}}, equivalently q|𝒰,a<n=p|𝒰,a<nq|_{\mathcal{U},a_{<n}}=p|_{\mathcal{U},a_{<n}}.

Suppose for a contradiction that φ(x,a<n)q|𝒰,a<n\varphi(x,a_{<n})\in q|_{\mathcal{U},a_{<n}} but ¬φ(x,a<n)p|𝒰,a<n\neg\varphi(x,a_{<n})\in p|_{\mathcal{U},a_{<n}}. Extend a<na_{<n} to a Morley sequence in pp by bnp|𝒰,a<n,bn+1p|𝒰,a<n,bn,b_{n}\models p|_{\mathcal{U},a_{<n}},b_{n+1}\models p|_{\mathcal{U},a_{<n},b_{n}},\dots. Since a<nbna_{<n}b_{\geq n} is a Morley sequence in pp, we get, by generic stability of pp, that the formula φ(x,a<n1,an1)¬φ(x,a<n1,bm)\varphi(x,a_{<n-1},a_{n-1})\wedge\neg\varphi(x,a_{<n-1},b_{m}) divides over 𝒰\mathcal{U} for every mnm\geq n. As by assumption we know that q|𝒰,a<n,bnq|_{\mathcal{U},a_{<n},b_{\geq n}} does not fork over 𝒰\mathcal{U}, we conclude that φ(x,a<n1,bm)q(x)\varphi(x,a_{<n-1},b_{m})\in q(x) for all mnm\geq n.

Pick any cq|𝒰,a<n,bnc\models q|_{\mathcal{U},a_{<n},b_{\geq n}}. By the above conclusion, φ(c,a<n1,bm)\models\varphi(c,a_{<n-1},b_{m}) for all mnm\geq n, so, by generic stability of pp, φ(c,a<n1,y)p(y)\varphi(c,a_{<n-1},y)\in p^{\prime}(y).

On the other hand, since ¬φ(x,a<n)p|𝒰,a<n\neg\varphi(x,a_{<n})\in p|_{\mathcal{U},a_{<n}} and bnb_{\geq n} is Morley sequence in pp over 𝒰,a<n\mathcal{U},a_{<n}, we get that ¬φ(bi,a<n)\models\neg\varphi(b_{i},a_{<n}) for all ini\geq n. Since a<nbma_{<n}b_{\geq m} is a Morley sequence in pp, and as such it is totally indiscernible, we get that ¬φ(an1,a<n1,bi)\models\neg\varphi(a_{n-1},a_{<n-1},b_{i}) for all ini\geq n. Therefore, ¬φ(an1,a<n1,y)p(y)\neg\varphi(a_{n-1},a_{<n-1},y)\in p^{\prime}(y).

Summarizing, the last two paragraphs yield φ(c,a<n1,y)¬φ(an1,a<n1,y)p(y)\varphi(c,a_{<n-1},y)\wedge\neg\varphi(a_{n-1},a_{<n-1},y)\in p^{\prime}(y). This contradicts the 𝒰\mathcal{U}-invariance of pp^{\prime}, because c𝒰,a<n1an1c\equiv_{\mathcal{U},a_{<n-1}}a_{n-1} as both these elements satisfy q|𝒰,a<n1q|_{\mathcal{U},a_{<n-1}}. ∎

Now consider any φ(x,b)q\varphi(x,b)\in q. Pick aiq|𝒰,a,b,a<ia_{i}\models q|_{\mathcal{U},a,b,a_{<i}} in 𝒰\mathcal{U}^{\prime} for all 0<i<ω0<i<\omega, and put a0:=aa_{0}:=a. By the claim, (ai)i<ω(a_{i})_{i<\omega} is a Morley sequence in pp. Since φ(ai,b)\models\varphi(a_{i},b) for all 0<i<ω0<i<\omega, by generic stability of pp, we get φ(x,b)p\varphi(x,b)\in p^{\prime}. Thus, q=pq=p^{\prime}. ∎

Now, to prove Proposition 2.29 via Method 2, using Fact 2.30, we get that RΔφ(p|𝒰B)=RΔφ(p)R_{\Delta_{\varphi}}(p^{\prime}|_{\mathcal{U}B})=R_{\Delta_{\varphi}}(p) for every set BB of realizations of pp (where p:=p|𝒰p^{\prime}:=p|_{\mathcal{U}^{\prime}}). This implies (by Fact 2.26(3) — which does not require any assumptions) that RΔφ((ap)|𝒰B)=RΔφ(a(p|𝒰B))=RΔφ(p)R_{\Delta_{\varphi}}((a\cdot p^{\prime})|_{\mathcal{U}B})=R_{\Delta_{\varphi}}(a\cdot(p^{\prime}|_{\mathcal{U}B}))=R_{\Delta_{\varphi}}(p) for every set BB of realizations of pp containing aa. Since papp\subseteq a\cdot p^{\prime} (by the idempotence of pp), we conclude, using Fact 2.30, that (ap)|𝒰B(a\cdot p^{\prime})|_{\mathcal{U}B} does not fork over 𝒰\mathcal{U} for every set BB of realizations of pp containing aa. Hence, ap=pa\cdot p^{\prime}=p^{\prime} by Proposition 2.31.

2.9. Simple theories

Assume that TT is a simple theory, and as usual that pSG(𝒰)p\in S_{G}(\mathcal{U}) is idempotent and generically stable over M𝒰M\prec\mathcal{U}. Method 2 extends from stable to simple theories using stratified Shelah degrees in place of stratified local ranks. In simple theories, they are wrongly defined in Definition 4.1.4 of [Wag00] (as they are not left-invariant due to the lack of associativity outside G¯\bar{G}). A way to fix it is to use suitable φ(x,y)\varphi(x,y) (as in Definition 2.25) or to apply Definition 4.3.5 of [Wag00] (in the special case of type-definable rather than hyper-definable groups). In any case, by [Wag00], stratified Shelah degrees satisfy items (1)–(4) of Fact 2.26, so Method 2 applies directly and yields the following generalization of Proposition 2.28:

Proposition 2.32.

Let TT be simple, GG an \emptyset-type-definable group, and pSG(𝒰)p\in S_{G}(\mathcal{U}) idempotent and generically stable. Then pp is generically transitive.

2.10. Rosy theories

In the case of groups in rosy theories, again we can apply Method 2, using stratified local thorn ranks. They were defined and studied in [EKP08] (see [EKP08, Definition 1.13]) in the case of definable groups, and extend easily to type-definable groups (using φ0\varphi_{0} as in Definition 2.25). By [EKP08], stratified local thorn ranks satisfy items (1)–(4) of Fact 2.26, with thorn forking in place forking in item (4). However, [GOU13, Theorem 3.4] tells us that if a type qS(B)q\in S(B) is generically stable and ABA\subseteq B, then qq forks over AA if and only if qq thorn forks over AA. We have all the tools to prove the following generalization of Proposition 2.32 via Method 2.

Before its statement, let us first recall local thorn ranks and define stratified local thorn ranks þΦ,Θ,kG\th_{\Phi,\Theta,k}^{G}:

Definition 2.33.

For a finite set Φ\Phi of partitioned formulas with object variables xx and parameter variables yy, a finite set of formulas Θ\Theta in variables y,zy,z, and natural number k>0k>0, the þΦ,Θ,k\th_{\Phi,\Theta,k}-rank is the unique function from the collection of all consistent formulas with parameters to Ord{}\textrm{Ord}\cup\{\infty\} satisfying: þΦ,Θ,k(ψ)α+1\th_{\Phi,\Theta,k}(\psi)\geq\alpha+1 if and only if there is φΦ\varphi\in\Phi, some θ(y,z)Θ\theta(y,z)\in\Theta and parameter cc such that:

  1. (1)

    þΦ,Θ,k(ψ(x)φ(x,a))α\th_{\Phi,\Theta,k}(\psi(x)\land\varphi(x,a))\geq\alpha for infinitely many aθ(y,c)a\models\theta(y,c), and

  2. (2)

    {φ(x,a):aθ(y,c)}\left\{\varphi\left(x,a\right):a\models\theta(y,c)\right\} is kk-inconsistent.

Given a (partial) type π(x)\pi(x) closed under conjunction we define þΦ,Θ,k(π(x))\th_{\Phi,\Theta,k}(\pi(x)) to be the minimum of þΦ,Θ,k(ψ)\th_{\Phi,\Theta,k}(\psi) for ψπ(x)\psi\in\pi(x).

Definition 2.34.
  1. (1)

    For a formula φ(x,y)\varphi(x,y)\in\mathcal{L}, let φ~(x,t,y):=(z)(φ(z,y)φ0(z)x=tz)\tilde{\varphi}(x,t,y):=(\exists z)(\varphi(z,y)\wedge\varphi_{0}(z)\wedge x=t\cdot z), where φ0(x)\varphi_{0}(x) is chosen in Section 2.2.

  2. (2)

    For a finite set Φ\Phi of formulas in variables x,yx,y, put Φ~:={φ~(x,t,y):φ(x,y)Φ}\widetilde{\Phi}:=\{\tilde{\varphi}(x,t,y):\varphi(x,y)\in\Phi\}. For a finite set of formulas Θ\Theta in variables y,zy,z, put Θ(t,y;t,z):={θ(y,z)t=t:θΘ}\Theta^{*}(t,y;t^{\prime},z):=\{\theta(y,z)\land t=t^{\prime}:\theta\in\Theta\}.

  3. (3)

    The stratified þΦ,Θ,kG\th_{\Phi,\Theta,k}^{G}-rank is defined as the unique function satisfying:
    þΦ,Θ,kG(ψ)α+1\th_{\Phi,\Theta,k}^{G}(\psi)\geq\alpha+1 if and only if there is a formula φΦ\varphi\in\Phi, some θ(t,y;tz)Θ\theta^{*}(t,y;t^{\prime}z)\in\Theta^{*} and parameters gGg\in G and cc anywhere such that:

    1. (a)

      þΦ,Θ,kG(ψ(x)φ~(x,g,b))n\th_{\Phi,\Theta,k}^{G}(\psi(x)\land\tilde{\varphi}(x,g,b))\geq n for infinitely many (g,b)θ(t,y;g,c)(g,b)\models\theta^{*}(t,y;g,c), and

    2. (b)

      {φ~(x,g,b):(g,b)θ(t,y;g,c)}\left\{\tilde{\varphi}(x,g,b):(g,b)\models\theta^{*}(t,y;g,c)\right\} is kk-inconsistent.

    Given a (partial) type π(x)\pi(x) closed under conjunction we define þΦ,Θ,kG(π(x))\th_{\Phi,\Theta,k}^{G}(\pi(x)) to be the minimum of þΦ,Θ,kG(ψ)\th_{\Phi,\Theta,k}^{G}(\psi) for ψπ(x)\psi\in\pi(x).

Proposition 2.35.

If TT is rosy and GG is a type-definable group, then every generically stable idempotent type pSG(𝒰)p\in S_{G}(\mathcal{U}) is generically transitive.

Proof.

Assume pp is generically stable over M𝒰M\prec\mathcal{U} and let apa\models p in 𝒰𝒰\mathcal{U}^{\prime}\succ\mathcal{U} and p=p|𝒰p^{\prime}=p|_{\mathcal{U}^{\prime}}. As usual, by Remark 2.12 it suffices to show that aStab(p)a\in\operatorname{Stab}_{\ell}(p^{\prime}). Since pp^{\prime} does not fork over 𝒰\mathcal{U}, it does not þ\th-fork over 𝒰\mathcal{U}, so þΦ,Θ,kG(p)=þΦ,Θ,kG(p)\th^{G}_{\Phi,\Theta,k}(p^{\prime})=\th^{G}_{\Phi,\Theta,k}(p) for every finite Φ\Phi, Θ\Theta, kk. Since þΦ,Θ,kG(ap)=þΦ,Θ,kG(p)\th^{G}_{\Phi,\Theta,k}(a\cdot p^{\prime})=\th^{G}_{\Phi,\Theta,k}(p^{\prime}), we conclude that þΦ,Θ,kG(ap)=þΦ,Θ,kG(p)\th^{G}_{\Phi,\Theta,k}(a\cdot p^{\prime})=\th^{G}_{\Phi,\Theta,k}(p). On the other hand, since pp is a generically stable idempotent and ap(𝒰)a\in p(\mathcal{U}^{\prime}), we have papp\subseteq a\cdot p^{\prime}. Therefore, apa\cdot p^{\prime} does not þ\th-fork over 𝒰\mathcal{U}. In order to conclude that apa\cdot p^{\prime} does not fork over 𝒰\mathcal{U} (and so coincides with pp^{\prime}, which finishes the proof), by virtue of [GOU13, Theorem 3.4], it remains to show the following:

Claim.

apa\cdot p^{\prime} is generically stable over M,aM,a.

The claim follows from Fact 2.3 since pp^{\prime} is generically stable over MM (and so over M,aM,a) and apa\cdot p^{\prime} is realized by agdcl(a,g)a\cdot g\in\operatorname{dcl}(a,g) for gpg\models p^{\prime}. ∎

2.11. Generic transitivity of pp and stratified rank localized on pp

As we saw in Sections 2.72.10, generic transitivity of an idempotent generically stable type pSG(𝒰)p\in S_{G}(\mathcal{U}) can be established using a well-behaved stratified rank. In this section, working in an arbitrary theory, we define an analog of the stratified rank in stable theories (Definition 2.25) restricting to formulas with parameters from a Morley sequence in a generically stable type pp (aiming for it to satisfy the properties similar to Fact 2.26 needed to apply Method 1 or 2 from the proof of Proposition 2.28). For technical reasons, we will make a stronger assumption that p(n)p^{(n)} is generically stable for all nn. In NIP theories, or even in NTP2 theories (by [CGH23]), this follows from generic stability of pp; but it is open in general if generic stability of pp implies that p(2)p^{(2)} is generically stable (a counterexample was suggested in [ACP14, Example 1.7], however it does not work — see [CGH23a, Section 8.1]). As the main result of this section, we will show that this rank is left invariant (under multiplication by elements from p(𝒰)p(\mathcal{U}^{\prime})) if and only if pp is generically transitive.

The following proposition will be used to show that our ranks are finite, but it may be of independent interest.

Proposition 2.36.

Let pS(𝒰)p\in S(\mathcal{U}), M𝒰M\prec\mathcal{U} a small model, and assume that p(n)p^{(n)} is generically stable over MM for every n>0n\in\mathbb{N}_{>0}. Let A𝒰A\subseteq\mathcal{U} be finite and let (ai)i<ω𝒰(a_{i})_{i<\omega}\subseteq\mathcal{U} be a Morley sequence in pp over MM. Let φ(x,y)(𝒰)\varphi(x,y)\in\mathcal{L}(\mathcal{U}) be any formula (possibly with parameters, and x,yx,y arbitrary tuples of variables). Then there are only countably many types in Sφ(A(ai)i<ω)S_{\varphi}(A(a_{i})_{i<\omega}).

Proof.

As there are only finitely many possibilities for substitutions of the elements of the finite set AA in place of some variables in yy, without loss of generality we may assume that A=A=\emptyset.

Denote n:=|y|n:=|y|, say y=(y0,,yn1)y=(y_{0},\ldots,y_{n-1}) with yjy_{j} corresponding to the free variables of pp for j<nj<n. The next claim follows from generic stability of p(n)p^{(n)}.

Claim.

There exists NφωN^{\varphi}\in\omega such that: for every b𝒰xb\in\mathcal{U}^{x} there is a subset SbφωS^{\varphi}_{b}\subseteq\omega of cardinality Nφ\leq N^{\varphi} and εbφ{0,1}\varepsilon^{\varphi}_{b}\in\{0,1\} such that for every pairwise distinct i0,,in1ωSbφi_{0},\dots,i_{n-1}\in\omega\setminus S^{\varphi}_{b} we have φ(b,ai0,,ain1)εbφ\models\varphi(b,a_{i_{0}},\dots,a_{i_{n-1}})^{\varepsilon^{\varphi}_{b}}.

Proof.

By generic stability of p(n)p^{(n)}, there is NφN_{\varphi} such that for every Morley sequence (c¯i)i<2Nφ(\bar{c}_{i})_{i<2N_{\varphi}} in p(n)p^{(n)} over MM and for every b𝒰xb\in\mathcal{U}^{x}, either for all but Nφ1N_{\varphi}-1 many ii’s we have φ(b,ci)\models\varphi(b,c_{i}), or for all but Nφ1N_{\varphi}-1 many ii’s we have ¬φ(b,ci)\models\neg\varphi(b,c_{i}). Put Nφ:=2nNφN^{\varphi}:=2nN_{\varphi}, suppose that it does not satisfy the requirement in the claim, and choose a witness b𝒰xb\in\mathcal{U}^{x} for that. Then, by recursion on kk, we can find pairwise distinct numbers ilki_{l}^{k} and jlkj_{l}^{k} for l<nl<n and k<Nφk<N_{\varphi} so that φ(b,ai0k,,ain1k)\models\varphi\left(b,a_{i_{0}^{k}},\dots,a_{i_{n-1}^{k}}\right) and ¬φ(b,aj0k,,ajn1k)\models\neg\varphi\left(b,a_{j_{0}^{k}},\dots,a_{j_{n-1}^{k}}\right) for all k<Nφk<N_{\varphi}. As

(ail0)l<n,(ajl0)l<n,,(ailNφ1)l<n,(ajlNφ1)l<n\left\langle\left(a_{i_{l}^{0}}\right)_{l<n},\left(a_{j_{l}^{0}}\right)_{l<n},\dots,\left(a_{i_{l}^{N_{\varphi}-1}}\right)_{l<n},\left(a_{j_{l}^{N_{\varphi}-1}}\right)_{l<n}\right\rangle

is a Morley sequence in p(n)p^{(n)} over MM, the previous sentence contradicts the choice of NφN_{\varphi}. ∎

By the claim, varying b𝒰xb\in\mathcal{U}^{x}, we have only countably many possibilities for the finite set SbφωS^{\varphi}_{b}\subset\omega and two possibilities for εbφ\varepsilon^{\varphi}_{b}, so only countably many possibilities for (Sbφ,εbφ)\left(S^{\varphi}_{b},\varepsilon^{\varphi}_{b}\right).

For j<nj<n, let φj(x,yj;y{yj}):=φ(x;y)\varphi_{j}(x,y_{j};y\setminus\{y_{j}\}):=\varphi(x;y). The proof of the proportion is by induction on n(=|y|)n(=|y|).

Base step (n=1n=1). For any b𝒰xb\in\mathcal{U}^{x}, the type tpφ(b/(ai)i<ω)\operatorname{tp}_{\varphi}(b/(a_{i})_{i<\omega}) is determined by the pair (Sbφ,εbφ)\left(S^{\varphi}_{b},\varepsilon^{\varphi}_{b}\right) and the truth values of finitely many sentences φ(b,ai)\varphi(b,a_{i}) for iSbφi\in S^{\varphi}_{b}. So we get countably many possibilities for tp(b/(ai)i<ω)\operatorname{tp}(b/(a_{i})_{i<\omega}).

Induction step (nn+1n\to n+1). By induction hypothesis |Sφj((ai)i<ω)|0|S_{\varphi_{j}}((a_{i})_{i<\omega})|\leq\aleph_{0}, so

|Sφ(x;y0,,yj1,aij,yj+1,,yn1)((ai)i<ω)|0\left\lvert S_{\varphi(x;y_{0},\dots,y_{j-1},a_{i_{j}},y_{j+1},\dots,y_{n-1})}((a_{i})_{i<\omega})\right\rvert\leq\aleph_{0}

for any j<nj<n and ijωi_{j}\in\omega. Since tpφ(b/(ai)i<ω)\operatorname{tp}_{\varphi}(b/(a_{i})_{i<\omega}) is determined by the pair (Sbφ,εbφ)(S^{\varphi}_{b},\varepsilon^{\varphi}_{b}) together with

j<nijSbφtpφ(x;y0,,yj1,aij,yj+1,,yn1)(b/(ai)i<ω),\bigcup_{j<n}\bigcup_{i_{j}\in S^{\varphi}_{b}}\operatorname{tp}_{\varphi(x;y_{0},\dots,y_{j-1},a_{i_{j}},y_{j+1},\dots,y_{n-1})}(b/(a_{i})_{i<\omega}),

we conclude that there are countably many types in Sφ((Ai)i<ω)S_{\varphi}((A_{i})_{i<\omega}). ∎

Now let G(x)G(x) be an \emptyset-type-definable group and pSG(𝒰)p\in S_{G}(\mathcal{U}) such that p(n)p^{(n)} is generically stable over M𝒰M\prec\mathcal{U} for all nωn\in\omega. Let 𝒰𝒰\mathcal{U}^{\prime}\succ\mathcal{U} be a bigger monster model. We define a version of stratified local ranks, where the inconsistent types witnessing the increase in rank have to be defined over a Morley sequence in pp.

Definition 2.37.
  1. (1)

    For a formula φ(x,y)\varphi(x,y)\in\mathcal{L}, let φ~(x,t,y):=(z)(φ(z,y)φ0(z)φ0(t)x=tz)\tilde{\varphi}(x,t,y):=(\exists z)(\varphi(z,y)\wedge\varphi_{0}(z)\wedge\varphi_{0}(t)\wedge x=t\cdot z), where φ0(x)\varphi_{0}(x) is chosen in Section 2.2.

  2. (2)

    Following Shelah’s terminology (rather than Pillay’s from [Pil96]), given A𝒰A\subseteq\mathcal{U}, by a φ~(x,t,y)\tilde{\varphi}(x,t,y)-formula over AA we mean a formula of the form φ~(x,a,b)\tilde{\varphi}(x,a,b) or ¬φ~(x,a,b)\neg\tilde{\varphi}(x,a,b) for a,ba,b from AA.

  3. (3)

    By a φ~(x,t,y)\tilde{\varphi}(x,t,y)-type over AA we mean a consistent collection of φ~(x,t,y)\tilde{\varphi}(x,t,y)-formulas over AA. Two such types are explicitly contradictory if there is a φ~(x,t,y)\tilde{\varphi}(x,t,y)-formula contained in one of these types such that the negation of this formula is in the other type.

Definition 2.38.
  1. (1)

    We define a function Rp,φR_{p,\varphi} from the collection of all partial types in xx over 𝒰\mathcal{U}^{\prime} to Ord{}\textrm{Ord}\cup\{\infty\} as a unique function satisfying: Rp,φ(π(x))α+1R_{p,\varphi}(\pi(x))\geq\alpha+1 if and only if for every finite π0(x)π(x)\pi_{0}(x)\subseteq\pi(x) and nωn\in\omega there exist pairwise explicitly contradictory φ~(x,t,y)\tilde{\varphi}(x,t,y)-types q0(x),,qn1(x)q_{0}(x),\dots,q_{n-1}(x) whose parameters altogether form a Morely sequence in pp over MM together with the parameters of π0(x)\pi_{0}(x) and such that Rp,φ(π0(x)qi(x))αR_{p,\varphi}(\pi_{0}(x)\cup q_{i}(x))\geq\alpha for all i<ni<n.

  2. (2)

    If Rp,φ(π(x))<R_{p,\varphi}(\pi(x))<\infty, then Mltp,φ(π(x))\operatorname{{Mlt}}_{p,\varphi}(\pi(x)) is defined as the maximal number nωn\in\omega such that for every finite π0(x)π(x)\pi_{0}(x)\subseteq\pi(x) with Rp,φ(π0(x))=Rp,φ(π(x))R_{p,\varphi}(\pi_{0}(x))=R_{p,\varphi}(\pi(x)) there are pairwise explicitly contradictory φ~(x,t,y)\tilde{\varphi}(x,t,y)-types q0(x),,qn1(x)q_{0}(x),\dots,q_{n-1}(x) whose parameters form a Morley sequence in pp over MM together with the parameters of π0(x)\pi_{0}(x) and such that Rp,φ(π0(x)qi(x))=Rp,φ(π(x))R_{p,\varphi}(\pi_{0}(x)\cup q_{i}(x))=R_{p,\varphi}(\pi(x)) for all i<ni<n.

Lemma 2.39.

Assume that p(n)p^{(n)} is generically stable for all nωn\in\omega. Then the ranks Rp,φR_{p,\varphi} have the following properties.

  1. (1)

    Rp,φ(x=x)<ωR_{p,\varphi}(x=x)<\omega.

  2. (2)

    Rp,φ(ψ1(x)ψ2(x))=max{Rp,φ(ψ1(x)),Rp,φ(ψ2(x))}R_{p,\varphi}(\psi_{1}(x)\lor\psi_{2}(x))=\max\left\{R_{p,\varphi}(\psi_{1}(x)),R_{p,\varphi}(\psi_{2}(x))\right\}.

  3. (3)

    Rp,φR_{p,\varphi} is invariant under automorphisms of 𝒰\mathcal{U}^{\prime} fixing MM pointwise.

  4. (4)

    In the definition of Rp,φR_{p,\varphi}, one can use π(x)\pi(x) in place of the finite piece π0(x)\pi_{0}(x).

  5. (5)

    Rp,φ(π(x))=Rp,φ(π(x)(z,t)(G(t)φ0(z)x=tz))R_{p,\varphi}(\pi(x))=R_{p,\varphi}(\pi(x)\cup(\exists z,t)(G(t)\wedge\varphi_{0}(z)\wedge x=t\cdot z)).

Proof.

(1) If not, then, by compactness, there is a tree (qη)η2<ω(q_{\eta})_{\eta\in 2^{<\omega}}, where each qηq_{\eta} is a φ~\tilde{\varphi}-formula, branches are consistent, the sons of every node are pairwise explicitly contradictory, and the parameters of the types qηq_{\eta} form a Morley sequence in pp over MM. So the number of complete φ~\tilde{\varphi}-types over these parameters is 202^{\aleph_{0}} which contradicts Proposition 2.36.

(2) and (3) are straightforward from the definitions.

(4) follows from (1) (namely, it is enough to consider only finite values αOrd\alpha\in\textrm{Ord} of the rank), and compactness. One just gets that Rp,φ(π(x))nR_{p,\varphi}(\pi(x))\geq n if and only if there is a tree (qη)ηnω(q_{\eta})_{\eta\in n^{\omega}} of φ~(x,t,y)\tilde{\varphi}(x,t,y)-types with suitable parameters and such that each branch is consistent with π(x)\pi(x) and the sons of every node are pairwise explicitly contradictory.

(5) follows from the observation that each but at most one of the types q0,,qn1q_{0},\dots,q_{n-1} in the definition of Rp,φR_{p,\varphi} implies (z,t)(G(t)φ0(z)x=tz)(\exists z,t)(G(t)\wedge\varphi_{0}(z)\wedge x=t\cdot z). ∎

We consider the left action of GG on partial types over 𝒰\mathcal{U}^{\prime} as follows:

Definition 2.40.

For gG(𝒰)g\in G(\mathcal{U}^{\prime}) and a partial type π(x)\pi(x) over a set A𝒰A\subseteq\mathcal{U}^{\prime}, we define gπ(x)g\cdot\pi(x) to be a partial type over A,gA,g defining the set g(π(𝒰)(G(𝒰)φ0(𝒰)))g\cdot(\pi(\mathcal{U}^{\prime})\cap(G(\mathcal{U}^{\prime})\cdot\varphi_{0}(\mathcal{U}^{\prime}))).

Proposition 2.41.

Generic transitivity of pp is equivalent to the invariance of the ranks Rp,φR_{p,\varphi} under action on the left by elements of p(𝒰)p(\mathcal{U}^{\prime}).

Proof.

()(\Rightarrow) By Lemma 2.39(5), it suffices to show that for any gp(𝒰)g\in p(\mathcal{U}^{\prime}) and any partial type π(x)\pi(x) which implies (y,z)(G(y)φ0(z)x=yz)(\exists y,z)(G(y)\wedge\varphi_{0}(z)\wedge x=yz) we have Rp,φ(gπ(x))=Rp,φ(π(x))R_{p,\varphi}(g\cdot\pi(x))=R_{p,\varphi}(\pi(x)).

By generic transitivity of pp and Remark 2.12 we have pSH(𝒰)p\in S_{H}(\mathcal{U}), where H=Stab(p)H=\operatorname{Stab}_{\ell}(p). Then the type-definable group HH is generically stable witnessed by pp, so p=p1p=p^{-1} by Fact 2.5. Hence, g1pg^{-1}\models p. Thus, since g1(gπ(x))g^{-1}\cdot(g\cdot\pi(x)) is equivalent to π(x)\pi(x) (by the choice of φ0(x)\varphi_{0}(x)), it is enough to show that Rp,φ(gπ(x))Rp,φ(π(x))R_{p,\varphi}(g\cdot\pi(x))\geq R_{p,\varphi}(\pi(x)). For that, by induction on α\alpha, we will show that Rp,φ(π(x))αR_{p,\varphi}(\pi(x))\geq\alpha implies that Rp,φ(gπ(x))αR_{p,\varphi}(g\cdot\pi(x))\geq\alpha.

The base step is obvious. For the induction step, consider any nωn\in\omega and pairwise explicitly contradictory φ~(x,t,y)\tilde{\varphi}(x,t,y)-types q0(x),,qn1(x)q_{0}(x),\dots,q_{n-1}(x) whose parameters form a Morley sequence (ci)i(c_{i})_{i} in pp over MM together with the parameters AA of π(x)\pi(x) and such that Rp,φ(π(x)qi(x))αR_{p,\varphi}(\pi(x)\cup q_{i}(x))\geq\alpha for all i<ni<n. Without loss of generality we may assume that this Morley sequence is over M,A,gM,A,g. By induction hypothesis, Rp,φ(g(π(x)qi(x)))αR_{p,\varphi}(g\cdot(\pi(x)\cup q_{i}(x)))\geq\alpha for all i<ni<n. On the other hand, by the choice of φ0(x)\varphi_{0}(x) and π(x)\pi(x), for any hG(𝒰)h\in G(\mathcal{U}^{\prime}) and any bb in 𝒰\mathcal{U}^{\prime} we have gφ~(x,h,b)g\cdot\tilde{\varphi}(x,h,b) is equivalent to φ~(x,gh,b)\tilde{\varphi}(x,g\cdot h,b) and g(π(x)¬φ~(x,h,b))g\cdot(\pi(x)\wedge\neg\tilde{\varphi}(x,h,b)) is equivalent to gπ(x)¬φ~(x,gh,b)g\cdot\pi(x)\wedge\neg\tilde{\varphi}(x,g\cdot h,b). Thus, Rp,φ(gπ(x)qi(x))αR_{p,\varphi}(g\cdot\pi(x)\cup q_{i}^{\prime}(x))\geq\alpha for all i<ni<n for some pairwise explicitly contradictory φ~(x,t,y)\tilde{\varphi}(x,t,y)-types q0(x),,qn1(x)q_{0}^{\prime}(x),\dots,q_{n-1}^{\prime}(x) whose parameters form a sequence (ci)i(c_{i}^{\prime})_{i}, where for every ii, ci=cic_{i}^{\prime}=c_{i} or ci=gcic_{i}^{\prime}=g\cdot c_{i}. It remains to show that (ci)i(c_{i}^{\prime})_{i} is a Morley sequence in pp over M,A,gM,A,g. So we need to show that cip|M,A,g,(cj)jic_{i}^{\prime}\models p|_{M,A,g,(c_{j}^{\prime})_{j\neq i}}.

If ci=cic_{i}^{\prime}=c_{i}, then this follows from the fact that cip|M,A,g,(cj)jic_{i}\models p|_{M,A,g,(c_{j})_{j\neq i}}. Consider the case ci=gcic_{i}^{\prime}=g\cdot c_{i}. Since cip|M,A,g,(cj)jic_{i}\models p|_{M,A,g,(c_{j})_{j\neq i}}, we have ci|MM,A,g,(cj)jic_{i}\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{M}M,A,g,(c_{j}^{\prime})_{j\neq i} and so gci|M,gM,A,g,(cj)jig\cdot c_{i}\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{M,g}M,A,g,(c_{j}^{\prime})_{j\neq i} by left transitivity. On the other hand, by generic transitivity of pp and Remark 2.12, gcip|M,gg\cdot c_{i}\models p|_{M,g}. Hence, ci=gcip|M,A,g,(cj)jic_{i}^{\prime}=g\cdot c_{i}\models p|_{M,A,g,(c_{j}^{\prime})_{j\neq i}} (by Fact 2.1(2)).

()(\Leftarrow) We will apply Method 1 from the proof of Proposition 2.28, using Proposition 2.31.

Let apa\models p in 𝒰\mathcal{U}^{\prime}, and suppose for a contradiction that appa\cdot p^{\prime}\neq p^{\prime}. By Proposition 2.31, there is a formula φ(x,a¯)\varphi(x,\bar{a}) with a¯\bar{a} being a Morley sequence in pp (over 𝒰\mathcal{U}) such that φ(x,a¯)p\varphi(x,\bar{a})\in p^{\prime} but ¬φ(x,a¯)ap\neg\varphi(x,\bar{a})\in a\cdot p^{\prime}. Then we follow the lines of Method 1, using Rp,φR_{p,\varphi} in place of RΔφR_{\Delta_{\varphi}} and left invariance of RΔp,φR_{\Delta_{p,\varphi}} under aa (which follows by assumption) together with invariance under Aut(𝒰/M)\operatorname{Aut}(\mathcal{U}^{\prime}/M). At the end, we get ppapp\subseteq p^{\prime}\cap a\cdot p^{\prime}, Rp,φ(p)=Rp,φ(p)=Rp,φ(ap)R_{p,\varphi}(p)=R_{p,\varphi}(p^{\prime})=R_{p,\varphi}(a\cdot p^{\prime}) and Mltp,φ(x,y¯)(p)=1\operatorname{{Mlt}}_{p,\varphi(x,\bar{y})}(p)=1, which contradicts the choice of φ(x,a¯)\varphi(x,\bar{a}). ∎

Problem 2.42.

Does the equivalence in Proposition 2.41 hold only assuming that pp is generically stable? (As opposed to p(n)p^{(n)} is generically stable for all nωn\in\omega.)

2.12. Stabilizer of a generically transitive type is an intersection of definable groups

As before, let GG be a type-definable group and let pSG(𝒰)p\in S_{G}(\mathcal{U}) be a generically stable type over MM. By Proposition 2.7, we know that Stab(p)\operatorname{Stab}_{\ell}(p) is an intersection of relatively MM-definable subgroups of G¯\bar{G}.

Question 2.43.

Assuming that pp is generically stable and idempotent, is Stab(p)\operatorname{Stab}_{\ell}(p) an intersection of MM-definable groups?

This question is open only in the case of type-definable G¯\bar{G} (in the definable case, the answer is trivially positive by the above comment).

Proposition 2.44.

If GG is type-definable and pSG(𝒰)p\in S_{G}(\mathcal{U}) is generically stable, idempotent and generically transitive, then the answer to Question 2.43 is positive.

Proof.

This proof is an elaboration on the proof of Hrushovski’s theorem that a type-definable group in a stable theory is an intersection of definable groups [Hru90].

Choose φ0(x)\varphi_{0}(x)\in\mathcal{L} as in Section 2.2, we will only use that G(x)φ0(x)G(x)\vdash\varphi_{0}(x) and \cdot is defined and associative on φ0(𝒰)\varphi_{0}(\mathcal{U}) and ae=a=eaa\cdot e=a=e\cdot a for all aφ0(𝒰)a\in\varphi_{0}(\mathcal{U}).

We will prove that there exists an MM-definable set H=Hφ0H=H_{\varphi_{0}} such that p|M(𝒰)Hφ0(𝒰)p|_{M}(\mathcal{U})\subseteq H\subseteq\varphi_{0}(\mathcal{U}) and (H,)(H,\cdot) is a group. Then Stab(p)Hφ0\operatorname{Stab}_{\ell}(p)\leq H_{\varphi_{0}} and φ0Hφ0G¯\bigcap_{\varphi_{0}}H_{\varphi_{0}}\leq\bar{G}, where φ0\varphi_{0} ranges over the formulas chosen as above. Therefore, by Proposition 2.7 and compactness, Stab(p)\operatorname{Stab}_{\ell}(p) is an intersection of MM-definable groups (all with group operation given by \cdot), so the proof will be finished.

We can clearly assume that the partial type G(x)G(x) is closed under conjunction. For any φ(x)G(x)\varphi(x)\in G(x) put δφ(x,y):=φ0(x)φ0(y)φ(yx)\delta_{\varphi}(x,y):=\varphi_{0}(x)\wedge\varphi_{0}(y)\wedge\varphi(y\cdot x). Let εφ(y)\varepsilon_{\varphi}(y) be the δφ(x;y)\delta_{\varphi}(x;y)-definition of pp; this is a formula over MM, since pp is definable over MM. Also, εφ(y)\varepsilon_{\varphi}(y) implies φ0(y)\varphi_{0}(y).

Claim 1.

G(x){εφ(x):φ(x)G(x)}G(x)\equiv\{\varepsilon_{\varphi}(x):\varphi(x)\in G(x)\}.

Proof.

()(\vdash) Consider any 𝒰bG(x)\mathcal{U}\ni b\models G(x), i.e. bG¯b\in\bar{G}. Take apa\models p. Then baG(𝒰)b\cdot a\in G(\mathcal{U}^{\prime}). Hence, for any φ(x)G(x)\varphi(x)\in G(x) we have φ0(a)φ0(b)φ(ba)\models\varphi_{0}(a)\wedge\varphi_{0}(b)\wedge\varphi(b\cdot a), and so bεφ(x)b\models\varepsilon_{\varphi}(x).

()(\dashv) Consider any 𝒰b{εφ(x):φ(x)G(x)}\mathcal{U}\ni b\models\{\varepsilon_{\varphi}(x):\varphi(x)\in G(x)\}. Take apa\models p. Then for every φ(x)G(x)\varphi(x)\in G(x) we have δφ(a,b)\models\delta_{\varphi}(a,b), so baφ(x)b\cdot a\models\varphi(x). Hence, baG(𝒰)b\cdot a\in G(\mathcal{U}^{\prime}), so b=(ba)a1G(𝒰)b=(b\cdot a)\cdot a^{-1}\in G(\mathcal{U}^{\prime}) (the equality holds, as bφ0(x)b\models\varphi_{0}(x) and aG(𝒰)a\in G(\mathcal{U}^{\prime})), whence bG(x)b\models G(x). ∎

By Claim 1, choose φ(x)G(x)\varphi(x)\in G(x) such that abφ0(𝒰)a\cdot b\in\varphi_{0}(\mathcal{U}) for all a,bεφ(𝒰)a,b\in\varepsilon_{\varphi}(\mathcal{U}).

Claim 2.

For every bεφ(𝒰)b\in\varepsilon_{\varphi}(\mathcal{U}) and ap|M(𝒰)a\in p|_{M}(\mathcal{U}) we have baεφ(𝒰)b\cdot a\in\varepsilon_{\varphi}(\mathcal{U}).

Proof.

Consider any aa and bb as above. Take any cG¯c\in\bar{G} realizing p|M,a,bp|_{M,a,b}. By generic transitivity and Remark 2.12, acp|M,aa\cdot c\models p|_{M,a}, and so ac|MM,aa\cdot c\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{M}M,a. On the other hand, by left transitivity, ac|M,aM,a,ba\cdot c\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{M,a}M,a,b. Hence, ac|MM,a,ba\cdot c\mathrel{\raise 0.86108pt\hbox{\ooalign{$|$\cr\raise-3.87495pt\hbox{$\smile$}}}}_{M}M,a,b by transitivity of forking for generically stable types (Fact 2.1), and so acp|M,ba\cdot c\models p|_{M,b}. Therefore, since bεφ(𝒰)b\in\varepsilon_{\varphi}(\mathcal{U}), we get b(ac)φ(x)b\cdot(a\cdot c)\models\varphi(x). Since a,b,cφ0(𝒰)a,b,c\in\varphi_{0}(\mathcal{U}), this means that (ba)cφ(x)(b\cdot a)\cdot c\models\varphi(x). On the other hand, cp|M,bac\models p|_{M,b\cdot a}. Moreover, since bεφ(𝒰)b\in\varepsilon_{\varphi}(\mathcal{U}) and ap|M(𝒰)G¯εφ(𝒰)a\in p|_{M}(\mathcal{U})\subseteq\bar{G}\subseteq\varepsilon_{\varphi}(\mathcal{U}) (the last inclusion holds by Claim 1), we have baφ0(𝒰)b\cdot a\in\varphi_{0}(\mathcal{U}). So we conclude that baεφ(x)b\cdot a\models\varepsilon_{\varphi}(x). ∎

Put H0:=εφ(𝒰)H_{0}:=\varepsilon_{\varphi}(\mathcal{U}) and H1:={aH0:baH0 for all bH0}H_{1}:=\{a\in H_{0}:b\cdot a\in H_{0}\textrm{ for all }b\in H_{0}\}. By Claim 2, p|M(𝒰)H1p|_{M}(\mathcal{U})\subseteq H_{1}, and clearly eH1e\in H_{1}. Both H0H_{0} and H1H_{1} are MM-definable. Using the choice of φ0(x)\varphi_{0}(x) and the inclusions H1H0φ0(𝒰)H_{1}\subseteq H_{0}\subseteq\varphi_{0}(\mathcal{U}), one easily checks that H1H_{1} is closed under \cdot. Finally, put H=Hφ0:={aH1:ab=ba=e for some bH1}H=H_{\varphi_{0}}:=\{a\in H_{1}:a\cdot b=b\cdot a=e\textrm{ for some }b\in H_{1}\}. It is clearly MM-definable.

By generic transitivity of pp (and Remark 2.11), the type-definable group Stab(p)\operatorname{Stab}_{\ell}(p) is generically stable witnessed by pp (Definition 2.4), so p=p1p=p^{-1} by Fact 2.5. Therefore, as we have seen above that p|M(𝒰)H1p|_{M}(\mathcal{U})\subseteq H_{1}, we get that p|M(𝒰)Hp|_{M}(\mathcal{U})\subseteq H. Summarizing, p|M(𝒰)HH1H0φ0(𝒰)p|_{M}(\mathcal{U})\subseteq H\subseteq H_{1}\subseteq H_{0}\subseteq\varphi_{0}(\mathcal{U}).

One easily checks, using the choice of φ0(x)\varphi_{0}(x) and the inclusion H1φ0(x)H_{1}\subseteq\varphi_{0}(x), that HH is closed under \cdot. Since Hφ0(𝒰)H\subseteq\varphi_{0}(\mathcal{U}), we have that \cdot is associative on HH and eHe\in H is neutral in HH. Also, for any aHa\in H there is bH1b\in H_{1} with ab=ba=ea\cdot b=b\cdot a=e which implies that bHb\in H. Hence, (H,)(H,\cdot) is a group. ∎

Remark 2.45.

If we drop the assumption that pp is generically transitive, then the main difficulty in adaptation of the above proof is in Claim 2. Namely, by exactly the same method we only get “For every bεφ(𝒰)b\in\varepsilon_{\varphi}(\mathcal{U}) and ap|M,b(𝒰)a\in p|_{M,b}(\mathcal{U}) we have baεφ(𝒰)b\cdot a\in\varepsilon_{\varphi}(\mathcal{U})”.

2.13. Chain conditions for type-definable groups on a generically stable type

One approach towards Problem 2.10 is to adapt Newelski’s “2-step generation” theorem [New91, Theorem 2.3] that provides a positive answer for types in stable theories. One ingredient is the existence of a smallest type-definable group containing a given type in a stable theory. In this section we investigate this question (and corresponding chain conditions and connected components) for groups type-definable using parameters from a Morley sequence of a generically stable type.

The following is a generalization of [DK12, Lemma 2.1] to type-definable groups (combining it with [HP18, Lemma 2.1]).

Lemma 2.46.

Let GG be an \emptyset-type-definable group and p(y)SG(𝒰)p(y)\in S_{G}(\mathcal{U}) a global type generically stable over a small set A𝒰A\subset\mathcal{U}.

  1. (1)

    Assume that H(x,y,z)H(x,y,z) is a countable partial type over AA, and let κ:=220\kappa:=2^{2^{\aleph_{0}}}. Then for any linear order II, any tuple c𝒰zc\in\mathcal{U}^{z} and any b¯:=(bi:iI)p(I)|Ac\bar{b}:=\left(b_{i}:i\in I\right)\models p^{(I)}|_{Ac}, if H(𝒰,bi,c)G(𝒰)G(𝒰)H(\mathcal{U},b_{i},c)\cap G(\mathcal{U})\leq G(\mathcal{U}) for all iIi\in I, then for any JIJ\subseteq I with |J|=κ|J|=\kappa we have

    iIH(𝒰,bi,c)G(𝒰)=iJH(𝒰,bi,c)G(𝒰).\bigcap_{i\in I}H(\mathcal{U},b_{i},c)\cap G(\mathcal{U})=\bigcap_{i\in J}H(\mathcal{U},b_{i},c)\cap G(\mathcal{U}).
  2. (2)

    Assume that p(n)p^{(n)} is generically stable over AA for all nωn\in\omega. Let H(x;y¯,z)H(x;\bar{y},z) with y¯=(yα:α<ω)\bar{y}=(y_{\alpha}:\alpha<\omega) be a countable partial type over AA. Then κ:=220\kappa:=2^{2^{\aleph_{0}}} satisfies the following: for any linear order II, any tuple c𝒰zc\in\mathcal{U}^{z} and any b¯:=(bi:iI)p(I)|Ac\bar{b}:=\left(b_{i}:i\in I\right)\models p^{(I)}|_{Ac}, if H(𝒰,b¯i¯,c)G(𝒰)G(𝒰)H(\mathcal{U},\bar{b}_{\bar{i}},c)\cap G(\mathcal{U})\leq G(\mathcal{U}), where b¯i¯=(biα:α<ω)\bar{b}_{\bar{i}}=(b_{i_{\alpha}}:\alpha<\omega), for all i¯=(iα:α<ω)Iω\bar{i}=(i_{\alpha}:\alpha<\omega)\in I^{\omega}, then for any JIJ\subseteq I with |J|=κ|J|=\kappa we have

    i¯IωH(𝒰,b¯i¯,c)G(𝒰)=i¯JωH(𝒰,b¯i¯,c)G(𝒰).\bigcap_{\bar{i}\in I^{\omega}}H(\mathcal{U},\bar{b}_{\bar{i}},c)\cap G(\mathcal{U})=\bigcap_{\bar{i}\in J^{\omega}}H(\mathcal{U},\bar{b}_{\bar{i}},c)\cap G(\mathcal{U}).
  3. (3)

    Assume that p(n)p^{(n)} is generically stable over AA for all nωn\in\omega. Let \mathcal{F} be a family of subgroups of G(𝒰)G(\mathcal{U}) that is invariant under AA-automorphisms and closed under (possibly infinite) intersections and supergroups (so for example could take \mathcal{F} to be all subgroups of G(𝒰)G(\mathcal{U})).

    Then for any linear order II, any b¯:=(bi:iI)p(I)|A\bar{b}:=\left(b_{i}:i\in I\right)\models p^{(I)}|_{A} and any JI,|J|κ:=220J\subseteq I,|J|\geq\kappa:=2^{2^{\aleph_{0}}}, the intersection of all subgroups of G(𝒰)G(\mathcal{U}) from \mathcal{F} type-definable with parameters from Ab¯A\bar{b} is a subgroup of G(𝒰)G(\mathcal{U}) in \mathcal{F} relatively type-definable over A(bi:iJ)A(b_{i}:i\in J).

Proof.

(1) Without loss of generality we may clearly assume that |I|κ|I|\geq\kappa; and extending (bi:iI)(b_{i}:i\in I) to a longer Morley sequence if necessary, we may assume |I|>κ|I|>\kappa. First we show that there exists some JI,|J|κJ\subseteq I,|J|\leq\kappa with iIH(𝒰,bi,c)G(𝒰)=iJH(𝒰,bi,c)G(𝒰)\bigcap_{i\in I}H(\mathcal{U},b_{i},c)\cap G(\mathcal{U})=\bigcap_{i\in J}H(\mathcal{U},b_{i},c)\cap G(\mathcal{U}). Assume not, then by induction on α<κ+\alpha<\kappa^{+} we can choose gαG(𝒰)g_{\alpha}\in G(\mathcal{U}) and pairwise distinct iαIi_{\alpha}\in I so that

gαG(𝒰)(β<αH(𝒰,biβ,c))H(𝒰,biα,c).g_{\alpha}\in G(\mathcal{U})\cap\left(\bigcap_{\beta<\alpha}H(\mathcal{U},b_{i_{\beta}},c)\right)\setminus H(\mathcal{U},b_{i_{\alpha}},c).

Let A0AA_{0}\subseteq A with |A0|=0|A_{0}|=\aleph_{0} contain the parameters of HH, and let 0\mathcal{L}_{0} be a countable sublanguage of \mathcal{L} containing all of the formulas in HH. By the choice of κ\kappa, applying Erdős-Rado and passing to a subsequence of (gα,biα:α<κ)(g_{\alpha},b_{i_{\alpha}}:\alpha<\kappa) we may assume that the sequence (gα,biα:α<ω)(g_{\alpha},b_{i_{\alpha}}:\alpha<\omega) is “22-indiscernible” with respect to 0(A0c)\mathcal{L}_{0}(A_{0}c), i.e. tuples (gα,biα,gβ,biβ)(g_{\alpha},b_{i_{\alpha}},g_{\beta},b_{i_{\beta}}) have the same 0\mathcal{L}_{0}-type over A0cA_{0}c for all α<β<ω\alpha<\beta<\omega, and either iα<iβi_{\alpha}<i_{\beta} for all α<β<ω\alpha<\beta<\omega, or iα>iβi_{\alpha}>i_{\beta} for all α<β<ω\alpha<\beta<\omega. Assume we are in the former case (the latter case is similar).

First assume that gαH(𝒰,biβ,c)g_{\alpha}\notin H(\mathcal{U},b_{i_{\beta}},c) for some β>α\beta>\alpha in ω\omega, then ¬φ(gα,biβ,c)\models\neg\varphi(g_{\alpha},b_{i_{\beta}},c) for some φ(x,y,z)H(x,y,z)\varphi(x,y,z)\in H(x,y,z). Hence, by 22-indiscernibility, we have φ(gα,biβ,c)β<α\models\varphi(g_{\alpha},b_{i_{\beta}},c)\iff\beta<\alpha for all α,βω\alpha,\beta\in\omega. Taking αω\alpha\in\omega sufficiently large, this contradicts generic stability of pp.

Otherwise we have gαH(𝒰,biβ,c)αβg_{\alpha}\in H(\mathcal{U},b_{i_{\beta}},c)\iff\alpha\neq\beta, for all α,βω\alpha,\beta\in\omega. Note that for any h1,h2G(𝒰)H(𝒰,bi0,c)h_{1},h_{2}\in G(\mathcal{U})\cap H(\mathcal{U},b_{i_{0}},c), we have h1g0h2H(𝒰,bi0,c)h_{1}\cdot g_{0}\cdot h_{2}\notin H(\mathcal{U},b_{i_{0}},c) (as g0H(𝒰,bi0,c)g_{0}\notin H(\mathcal{U},b_{i_{0}},c)). By compactness there is a formula φ(x,y,z)H(x,y,z)\varphi(x,y,z)\in H(x,y,z) (without loss of generality the partial type HH is closed under conjunctions) so that: for any h1,h2G(𝒰)H(𝒢,bi0,c)h_{1},h_{2}\in G(\mathcal{U})\cap H(\mathcal{G},b_{i_{0}},c), ¬φ(h1g0h2,bi0,c)\models\neg\varphi(h_{1}\cdot g_{0}\cdot h_{2},b_{i_{0}},c). By 11-indiscernibility we then have: for any α<ω\alpha<\omega, for any h1,h2G(𝒰)H(𝒰,biα,c)h_{1},h_{2}\in G(\mathcal{U})\cap H(\mathcal{U},b_{i_{\alpha}},c), ¬φ(h1gαh2,biα,c)\models\neg\varphi(h_{1}\cdot g_{\alpha}\cdot h_{2},b_{i_{\alpha}},c). Let now KK be any finite subset of ω\omega, and let gK:=αKgiαg_{K}:=\prod_{\alpha\in K}g_{i_{\alpha}}. As gαH(𝒰,biβ,c)αβg_{\alpha}\in H(\mathcal{U},b_{i_{\beta}},c)\iff\alpha\neq\beta, it follows that for all α<ω\alpha<\omega, φ(gK,biα,c)αK\models\varphi(g_{K},b_{i_{\alpha}},c)\iff\alpha\notin K. Taking KK sufficiently large, this again contradicts generic stability of pp (in fact, even “generic NIP” of pp).

We have thus shown that there exists JI,|J|=κJ\subseteq I,|J|=\kappa with G(𝒰)iIH(𝒰,bi,c)=G(𝒰)iJH(𝒰,bi,c)G(\mathcal{U})\cap\bigcap_{i\in I}H(\mathcal{U},b_{i},c)=G(\mathcal{U})\cap\bigcap_{i\in J}H(\mathcal{U},b_{i},c). Given an arbitrary JI,|J|=κJ^{\prime}\subseteq I,|J^{\prime}|=\kappa, there exists a permutation σ\sigma of II sending JJ to JJ^{\prime}. As b¯\bar{b} is totally indiscernible over AcAc by generic stability of pp, there exists fAut(𝒰/Ac)f\in\operatorname{Aut}_{\mathcal{L}}(\mathcal{U}/Ac) with f(bi)=bσ(i)f(b_{i})=b_{\sigma(i)} for all ii. Applying ff we thus have:

G(𝒰)iIH(𝒰,bi,c)=G(𝒰)iIH(𝒰,bσ(i),c)\displaystyle G(\mathcal{U})\cap\bigcap_{i\in I}H(\mathcal{U},b_{i},c)=G(\mathcal{U})\cap\bigcap_{i\in I}H(\mathcal{U},b_{\sigma(i)},c)
=G(𝒰)iJH(𝒰,bσ(i),c)=G(𝒰)iJH(𝒰,bi,c).\displaystyle=G(\mathcal{U})\cap\bigcap_{i\in J}H(\mathcal{U},b_{\sigma(i)},c)=G(\mathcal{U})\cap\bigcap_{i\in J^{\prime}}H(\mathcal{U},b_{i},c).

(2) Let JI,|J|=κJ\subseteq I,|J|=\kappa be arbitrary. Fix any i¯=(iα:α<ω)Iω\bar{i}=(i_{\alpha}:\alpha<\omega)\in I^{\omega}. Let i¯:=(iα:α<ω,iαJ),i¯′′:=(iα:α<ω,iαJ)\bar{i}^{\prime}:=\left(i_{\alpha}:\alpha<\omega,i_{\alpha}\notin J\right),\bar{i}^{\prime\prime}:=\left(i_{\alpha}:\alpha<\omega,i_{\alpha}\in J\right), so that i¯=i¯i¯′′\bar{i}=\bar{i}^{\prime}\bar{i}^{\prime\prime} and i¯i¯′′=\bar{i}^{\prime}\cap\bar{i}^{\prime\prime}=\emptyset. As |i¯′′|0|\bar{i}^{\prime\prime}|\leq\aleph_{0}, we can choose tuples i¯αIω\bar{i}_{\alpha}\in I^{\leq\omega} for α<κ\alpha<\kappa so that: i¯0=i¯\bar{i}_{0}=\bar{i}^{\prime}, i¯αJω\bar{i}_{\alpha}\in J^{\leq\omega} for α>0\alpha>0, |i¯α|=|i¯||\bar{i}_{\alpha}|=|\bar{i}^{\prime}|, i¯αi¯′′=\bar{i}_{\alpha}\cap\bar{i}^{\prime\prime}=\emptyset and i¯αi¯β=\bar{i}_{\alpha}\cap\bar{i}_{\beta}=\emptyset for all αβ<κ\alpha\neq\beta<\kappa. As b¯p(I)|Ac\bar{b}\models p^{(I)}|_{Ac} is totally indiscernible over AcAc, it follows that (b¯i¯α:α<κ)(p|i¯|)Acb¯i¯′′κ\left(\bar{b}_{\bar{i}_{\alpha}}:\alpha<\kappa\right)\models\left(p^{|\bar{i}^{\prime}|}\right)^{\kappa}_{Ac\bar{b}_{\bar{i}^{\prime\prime}}}, hence applying (1) to the family H(x;b¯i¯α;cb¯i¯′′):=H(x;b¯i¯αb¯i¯′′;c)H^{\prime}\left(x;\bar{b}_{\bar{i}_{\alpha}};c\bar{b}_{\bar{i}^{\prime\prime}}\right):=H\left(x;\bar{b}_{\bar{i}_{\alpha}}\bar{b}_{\bar{i}^{\prime\prime}};c\right), α<κ\alpha<\kappa, and generically stable type p(ω)p^{(\omega)} we conclude that

G(𝒰)i¯JωH(𝒰,b¯i¯,c)G(𝒰)α<κ,α0H(𝒰;b¯i¯α;cb¯i¯′′)\displaystyle G(\mathcal{U})\cap\bigcap_{\bar{i}\in J^{\omega}}H(\mathcal{U},\bar{b}_{\bar{i}},c)\subseteq G(\mathcal{U})\cap\bigcap_{\alpha<\kappa,\alpha\neq 0}H^{\prime}\left(\mathcal{U};\bar{b}_{\bar{i}_{\alpha}};c\bar{b}_{\bar{i}^{\prime\prime}}\right)
G(𝒰)H(𝒰;b¯i¯0;cb¯i¯′′)=G(𝒰)H(𝒰,b¯i¯,c).\displaystyle\subseteq G(\mathcal{U})\cap H^{\prime}\left(\mathcal{U};\bar{b}_{\bar{i}_{0}};c\bar{b}_{\bar{i}^{\prime\prime}}\right)=G(\mathcal{U})\cap H(\mathcal{U},\bar{b}_{\bar{i}},c).

(3) Let HH\in\mathcal{F} be type-defined by a partial type π(x)\pi(x) with parameters in b¯A\bar{b}A for some b¯p(I)|A\bar{b}\models p^{(I)}|_{A} with II a small linear order. We can write H=G(𝒰)α<γHα(𝒰,b¯α)H=G(\mathcal{U})\cap\bigcap_{\alpha<\gamma}H_{\alpha}(\mathcal{U},\bar{b}_{\alpha}) for some ordinal γ\gamma, with each Hα(x,b¯α)π(x)H_{\alpha}(x,\bar{b}_{\alpha})\subseteq\pi(x) a countable partial type relatively defining a subgroup of G(𝒰)G(\mathcal{U}), and b¯α\bar{b}_{\alpha} a countable subsequence of b¯\bar{b}. In particular G(𝒰)Hα(𝒰,b¯α)G(\mathcal{U})\cap H_{\alpha}(\mathcal{U},\bar{b}_{\alpha})\in\mathcal{F}, as \mathcal{F} is closed under supergroups. Fix any JI,|J|=κJ\subseteq I,|J|=\kappa. For any α<γ\alpha<\gamma, applying (2) we have G(𝒰)j¯JωHα(𝒰,b¯i¯)G(𝒰)Hα(𝒰,b¯α)G(\mathcal{U})\cap\bigcap_{\bar{j}\in J^{\omega}}H_{\alpha}(\mathcal{U},\bar{b}_{\bar{i}})\subseteq G(\mathcal{U})\cap H_{\alpha}(\mathcal{U},\bar{b}_{\alpha}). And by indiscernibility of b¯\bar{b} over AA, AA-invariance of \mathcal{F} and closure under intersections, we have G(𝒰)j¯JωHα(𝒢,b¯i¯)G(\mathcal{U})\cap\bigcap_{\bar{j}\in J^{\omega}}H_{\alpha}(\mathcal{G},\bar{b}_{\bar{i}})\in\mathcal{F}. Hence G(𝒰)HG(𝒰)α<γ,j¯JωHα(𝒰,b¯i¯)G(\mathcal{U})\cap H\supseteq G(\mathcal{U})\cap\bigcap_{\alpha<\gamma,\bar{j}\in J^{\omega}}H_{\alpha}(\mathcal{U},\bar{b}_{\bar{i}})\in\mathcal{F}. ∎

Question 2.47.

Does Lemma 2.46(2) hold only assuming that pp is generically stable? The answer is positive for the analog for definable groups instead of type-definable groups ([DK12, Lemma 2.1]), but the proof there does not immediately seem to generalize to the type-definable case. We also expect that an analog of Lemma 2.46 holds for invariant instead of type-definable subgroups, but do not pursue it here.

Corollary 2.48.

Let GG be an \emptyset-type-definable group and pSG(𝒰)p\in S_{G}(\mathcal{U}) a global type so that p(n)p^{(n)} is generically stable for all n<ωn<\omega. Assume that pp is invariant over a small set A𝒰A\subset\mathcal{U}, and consider the family

p,A:={HG(𝒰):H is type-definable over Ab¯,b¯p(|b¯|)|A,pH(x)}.\mathcal{H}_{p,A}:=\left\{H\leq G(\mathcal{U}):H\textrm{ is type-definable over }A\bar{b},\bar{b}\models p^{(|\bar{b}|)}|_{A},p\vdash H(x)\right\}.

Then the group Gp,A:=p,AG_{p,A}:=\bigcap\mathcal{H}_{p,A} is type-definable over AA, and p|AGp,Ap|_{A}\vdash G_{p,A}.

Proof.

Let κ\kappa be as in Lemma 2.46, and fix some b¯p(κ)|A\bar{b}\models p^{(\kappa)}|_{A}. Assume we are given an arbitrary small linear order II and arbitrary b¯p(I)|A\bar{b}^{\prime}\models p^{(I)}|_{A}, and let HG(𝒰)H^{\prime}\leq G(\mathcal{U}) with pHp\vdash H^{\prime} be type-definable over b¯A\bar{b}^{\prime}A. We can choose some b¯′′p(κ)|Ab¯b¯\bar{b}^{\prime\prime}\models p^{(\kappa)}|_{A\bar{b}\bar{b}^{\prime}}. Let :={HG(𝒰):pH(x)}\mathcal{F}:=\left\{H\leq G(\mathcal{U}):p\vdash H(x)\right\}. Then the family \mathcal{F} is Aut(𝒰/A)\operatorname{Aut}(\mathcal{U}/A)-invariant by AA-invariance of pp, and is closed under supergroups and intersections. Applying Lemma 2.46(3) with \mathcal{F} and the sequence b¯+b¯′′p(I+κ)|A\bar{b}^{\prime}+\bar{b}^{\prime\prime}\models p^{(I+\kappa)}|_{A}, we find some H′′H^{\prime\prime}\in\mathcal{F} type-definable over b¯′′A\bar{b}^{\prime\prime}A with H′′HH^{\prime\prime}\subseteq H^{\prime}. Applying Lemma 2.46(3) again to the sequence b¯+b¯′′p(κ+κ)|A\bar{b}+\bar{b}^{\prime\prime}\models p^{(\kappa+\kappa)}|_{A}, we find some HH\in\mathcal{F} type-definable over b¯A\bar{b}A with HH′′H\subseteq H^{\prime\prime}. This shows that the group Gp,A:=p,AG_{p,A}:=\bigcap\mathcal{H}_{p,A} is type-definable, over Ab¯A\bar{b}. But it is also Aut(𝒰/A)\operatorname{Aut}(\mathcal{U}/A)-invariant (since the family p,A\mathcal{H}_{p,A} is Aut(𝒰/A)\operatorname{Aut}(\mathcal{U}/A)-invariant by AA-invariance of pp). Hence Gp,AG_{p,A} is type-definable over AA, and p|AGp,A(x)p|_{A}\vdash G_{p,A}(x). ∎

Question 2.49.
  1. (1)

    Does Gp,AG_{p,A} depend on AA?

  2. (2)

    Is Gp,AG_{p,A} an intersection of definable groups?

  3. (3)

    Let pSG(𝒰)p\in S_{G}(\mathcal{U}) be a generically stable type (idempotent and such that p(n)p^{(n)} is generically stable for all nωn\in\omega, if it helps). Does there exist a smallest type-definable (over a small set of parameters anywhere in 𝒰\mathcal{U}) group HH with pH(x)p\vdash H(x)?

All of these questions have a positive answer assuming pp is idempotent and generically transitive:

Proposition 2.50.

Assume that pSG(𝒰)p\in S_{G}(\mathcal{U}) is generically stable, idempotent and generically transitive. Then Stab(p)\operatorname{Stab}_{\ell}(p) is the smallest type-definable (with parameters anywhere in 𝒰\mathcal{U}) subgroup HH of GG so that pHp\vdash H.

Moreover, if p(n)p^{(n)} is generically stable for all nωn\in\omega, then Stab(p)=Gp,A\operatorname{Stab}_{\ell}(p)=G_{p,A} for any small A𝒰A\subseteq\mathcal{U} so that pp is invariant over AA, and Gp,AG_{p,A} is an intersection of definable groups.

Proof.

The first part is by Remark 2.11. Moreover, let A𝒰A\subseteq\mathcal{U} be such that pp is invariant over AA. As Stab(p)\operatorname{Stab}_{\ell}(p) is type-definable over AA (by Proposition 2.7) and pStab(p)p\vdash\operatorname{Stab}_{\ell}(p), by the definition of Gp,AG_{p,A} we have Gp,AStab(p)G_{p,A}\subseteq\operatorname{Stab}_{\ell}(p), and Stab(p)Gp,A\operatorname{Stab}_{\ell}(p)\subseteq G_{p,A} by minimality as Gp,AG_{p,A} is type-definable. Then Gp,AG_{p,A} is an intersection of definable groups by Proposition 2.44. ∎

We note however that we cannot answer Question 2.49(3) positively by proving a chain condition for groups containing pp and (type-)definable with arbitrary parameters:

Example 2.51.

Let 𝒰=(K,Γ,k)\mathcal{U}=(K,\Gamma,k) be a monster model of the theory ACVF of algebraically closed fields, and let G:=(K,+)G:=\left(K,+\right) be the additive group. For bK,βΓb\in K,\beta\in\Gamma and {,>}\square\in\{\geq,>\}, let Bβ(b):={xK:v(xb)β}B_{\square\beta}(b):=\{x\in K:v(x-b)\square\beta\} be the closed (respectively, open) ball of radius β\beta with center bb. Fix αΓ\alpha\in\Gamma, given a ball Bα(a)B_{\square\alpha}(a), by quantifier elimination in ACVF, density of Γ\Gamma, and the fact that for any two balls either one is contained in the other, or they are disjoint,

{xBα(a)}{xBβ(b):bK,βΓ,Bβ(b)Bα(a)}\left\{x\in B_{\square\alpha}(a)\right\}\cup\left\{x\notin B_{\square\beta}(b):b\in K,\beta\in\Gamma,B_{\square\beta}(b)\subsetneq B_{\square\alpha}(a)\right\}

determines a complete type pα,a(x)SG(𝒰)p_{\square\alpha,a}(x)\in S_{G}(\mathcal{U}), called the generic type of Bα(a)B_{\square\alpha}(a).

Now let B:=Bα(0)B:=B_{\geq\alpha}(0) and p:=pα,0p:=p_{\geq\alpha,0}. As for any a,ba,b we have a+Bβ(b)=Bβ(a+b)a+B_{\square\beta}(b)=B_{\square\beta}(a+b), hence for any bBb\in B we have b+B=Bb+B=B and if BBB^{\prime}\subsetneq B, then b+BBb+B^{\prime}\subseteq B for any ball BB^{\prime}. And if bKBb\in K\setminus B then (b+B)B=(b+B)\cap B=\emptyset, so we have Stab(p)=B\operatorname{Stab}_{\ell}(p)=B, and BG(𝒰)B\leq G(\mathcal{U}). In particular pBp\vdash B, and pp is left-BB-invariant. Note that pp generically stable over any aKa\in K with v(a)=αv(a)=\alpha (as it is internal to the residue field, or can see directly that it commutes with itself). Hence pp is idempotent.

Note that {B>β(0):βΓ,β<α}\{B_{>\beta}(0):\beta\in\Gamma,\beta<\alpha\} is a large strictly descending chain of definable subgroups of G(𝒰)G(\mathcal{U}) with pB>β(0)p\vdash B_{>\beta}(0) for all β<α\beta<\alpha. So we have arbitrary long descending chains of definable subgroups of G(𝒰)G(\mathcal{U}) containing pp (but the smallest one BB is still definable).

3. Idempotent generically stable measures

3.1. Overview

The main aim of this section is to prove the following:

Theorem.

(Theorem 3.45) Let G(x)G(x) be an abelian type-definable group, and assume that μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is fim (Definition 3.5) and idempotent. Then μ\mu is the unique left-invariant (and the unique right-invariant) measure on a type-definable subgroup of G(𝒰)G(\mathcal{U}) (namely, its stabilizer).

Remark 3.1.

In particular, if TT is NIP and GG is abelian, there is a one-to-one correspondence between generically stable idempotent measures and type-definable fsg subgroups of GG.

Remark 3.2.

The assumption that μ\mu is fim cannot be relaxed.

Indeed, consider G:=S1×G:=S^{1}\times\mathbb{R} as a definable group in (;+,×,0,1)(\mathbb{R};+,\times,0,1), and let G(𝒰)=𝒮1×G(\mathcal{U})=\mathcal{S}^{1}\times\mathcal{R} be a monster model. Let λ\lambda be the normalized Haar measure on S1S^{1} and pp the type of the cut above 0. Let λ\lambda^{\prime} be the unique smooth extension of λ\lambda to the monster model, pp^{\prime} the unique definable extension of pp, and μ:=λ×p\mu:=\lambda^{\prime}\times p^{\prime}. Then μ\mu is definable and idempotent. But Stab(μ)={(α,0):α𝒮1}\operatorname{Stab}(\mu)=\{(\alpha,0):\alpha\in\mathcal{S}^{1}\}, in particular μ([Stab(μ)])=0\mu([\operatorname{Stab}(\mu)])=0, so μ\mu cannot satisfy the conclusion of the theorem by Remark 3.39 (see also [CG23, Example 4.5]).

This section of the paper is organized as follows. We briefly recall the setting and some properties of Keisler measures in Section 3.2. In Section 3.3 we recall some basic facts and make some new observations involving fim measures — they provide a generalization of generically stable measures from NIP to arbitrary theories in the same way as generically stable types in the sense of [PT11] provide a generalization from NIP to arbitrary theories. In Section 3.4 we prove that the usual characterization of generic stability — any Morley sequence of a fim  measure determines the measure of arbitrary formulas by averaging along it — holds even when the parameters of these formulas are themselves replaced by a measure, see Theorem 3.13. In Section 3.5 we collect some basic facts about definable pushforwards of Keisler measures. In Section 3.6 we develop some theory of fim groups (i.e. groups admitting a translation invariant fim measure) in arbitrary theories, simultaneously generalizing from fsg groups in NIP theories and generically stable groups in arbitrary theories (in particular, that this translation invariant measure is unique and bi-invariant, Proposition 3.32). In Section 3.7 we develop an appropriate analog of generic transitivity for fim measures, generalizing some of the results for generically stable types from Section 2.4. Finally, in Section 3.8, we put all of this together in order to prove the main theorem of the section, adapting the weight argument from Section 2.5 to a purely measure theoretic context.

In Section 3.9 we isolate a weaker property of support transitivity and connect it to the algebraic properties of the semigroup induced by \ast on the support of an idempotent measure. In Section 3.10 we illustrate how the Keisler randomization can be used to reduce generic transitivity to support transitivity in stable groups, and discuss more general situations when the randomization may have an appropriate stratified rank.

3.2. Setting and notation

We work in the same setting as in Section 2.2. For a partitioned formula φ(x,y)\varphi(x,y), φ(y,x)\varphi^{*}(y,x) is the same formula but with the roles of the variables swapped. In the group setting, if φ(x,y)φ0(x)φ0(y)\varphi(x,y)\vdash\varphi_{0}(x)\land\varphi_{0}(y), then φ(x,y):=φ(xy)\varphi^{\prime}(x,y):=\varphi(x\cdot y). Let A𝒰A\subseteq\mathcal{U}. Then a Keisler measure (in variable xx over AA) is a finitely additive probability measure on x(A)\mathcal{L}_{x}(A) (modulo logical equivalence). We denote the collection of Keisler measures (in variable xx over AA) as 𝔐x(A)\mathfrak{M}_{x}(A). Given μ𝔐x(A)\mu\in\mathfrak{M}_{x}(A), we let S(μ)S(\mu) denote the support of μ\mu, i.e. the (closed) set of all pSx(A)p\in S_{x}(A) such that μ(φ(x))>0\mu(\varphi(x))>0 for every φ(x)p\varphi(x)\in p. We refer the reader to [CG22, CG23] for basic definitions involving Keisler measures (e.g. Borel-definable, definable, support of a measure, etc). Given a partial type π(x)\pi(x) over 𝒰\mathcal{U}, we will consider the closed set 𝔐π(𝒰):={μ𝔐x(𝒰):pS(μ)p(x)π(x)}\mathfrak{M}_{\pi}(\mathcal{U}):=\left\{\mu\in\mathfrak{M}_{x}(\mathcal{U}):p\in S(\mu)\Rightarrow p(x)\vdash\pi(x)\right\} of measures supported on π(x)\pi(x).

We will assume some familiarity with the basic theory of Keisler measures and their basic properties such as (automorphism-) invariance, (Borel-) definability, finite satisfiability, etc., and refer to e.g. [Sim15] or earlier papers in the series [CG22, CG23] for the details and references.

3.3. Fim measures

Throughout this section we work in an arbitrary theory TT, unless explicitly specified otherwise.

Definition 3.3.

Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}), ν𝔐y(𝒰)\nu\in\mathfrak{M}_{y}(\mathcal{U}) and suppose that μ\mu is Borel-definable. Then we define the Morley product of μ\mu and ν\nu, denoted μν\mu\otimes\nu, as the unique measure in 𝔐xy(𝒰)\mathfrak{M}_{xy}(\mathcal{U}) such that for any φ(x,y)xy(𝒰)\varphi(x,y)\in\mathcal{L}_{xy}(\mathcal{U}), we have

(μν)(φ(x,y))=Sy(A)Fμ,Aφd(ν|A^),(\mu\otimes\nu)(\varphi(x,y))=\int_{S_{y}(A)}F_{\mu,A}^{\varphi}d(\widehat{\nu|_{A}}),

where:

  1. (1)

    μ\mu is AA-invariant and AA contains all the parameters from φ\varphi,

  2. (2)

    Fμ,Aφ:Sy(A)[0,1]F_{\mu,A}^{\varphi}:S_{y}(A)\to[0,1] is defined by Fμ,Aφ(q)=μ(φ(x,b))F_{\mu,A}^{\varphi}(q)=\mu(\varphi(x,b)) for some (equivalently, any) bqb\models q in 𝒰\mathcal{U},

  3. (3)

    ν|A^\widehat{\nu|_{A}} is the unique regular Borel probability measure on Sx(A)S_{x}(A) corresponding to the Keisler measure ν|A\nu|_{A}.

See e.g. [CG22, Section 3.1] for an explanation why this product is well-defined and its basic properties. We will often abuse the notation slightly and replace ν|A^\widehat{\nu|_{A}} with either ν|A\nu|_{A} or simply ν\nu when it is clear from the context, and sometimes write Fμ,AφF_{\mu,A}^{\varphi} as FμφF_{\mu}^{\varphi}.

Definition 3.4.

Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) and suppose that μ\mu is Borel-definable. Then we define μ(1):=μ(x1)\mu^{(1)}:=\mu(x_{1}), μ(n+1)(x1,,xn+1):=μ(xn+1)μ(n)(x1,,xn)\mu^{(n+1)}(x_{1},\ldots,x_{n+1}):=\mu(x_{n+1})\otimes\mu^{(n)}(x_{1},\ldots,x_{n}), and μ(ω)=n<ωμ(n)(x1,,xn)\mu^{(\omega)}=\bigcup_{n<\omega}\mu^{(n)}(x_{1},\ldots,x_{n}). (In general, \otimes need not be commutative/associative on Borel definable measures in arbitrary theories.)

Definition 3.5.

[HPS13] Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) and M𝒰M\prec\mathcal{U} a small model. A Borel-definable measure μ\mu is fim (a frequency interpretation measure) over MM if μ\mu is MM-invariant and for any \mathcal{L}-formula φ(x,y)\varphi(x,y) there exists a sequence of formulas (θn(x1,,xn))1n<ω(\theta_{n}(x_{1},\ldots,x_{n}))_{1\leq n<\omega} in (M)\mathcal{L}(M) such that:

  1. (1)

    for any ε>0\varepsilon>0, there exists some nεωn_{\varepsilon}\in\omega satisfying: for any knεk\geq n_{\varepsilon}, if 𝒰θk(a¯)\mathcal{U}\models\theta_{k}(\bar{a}) then

    supb𝒰y|Av(a¯)(φ(x,b))μ(φ(x,b))|<ε;\sup_{b\in\mathcal{U}^{y}}|\operatorname{Av}(\bar{a})(\varphi(x,b))-\mu(\varphi(x,b))|<\varepsilon;
  2. (2)

    limnμ(n)(θn(x¯))=1\lim_{n\to\infty}\mu^{(n)}\left(\theta_{n}\left(\bar{x}\right)\right)=1.

We say that μ\mu is fim if μ\mu is fim over some small M𝒰M\prec\mathcal{U}.

Remark 3.6.

In NIP theories, fim is equivalent to each of the following two properties for measures: dfs (definable and finitely satisfiable) and fam (finitely approximable [CS21]), recovering the usual notion of generic stability for Keisler measures [HPS13]. Outside of the NIP context, fim (properly) implies fam over a model, which in turn (properly) implies dfs (see [CG20, CGH23a]).

Generalizing from generically stable measures in NIP, one has:

Fact 3.7.

[CGH23a] If μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) is fim and ν𝔐y(𝒰)\nu\in\mathfrak{M}_{y}(\mathcal{U}) is Borel definable, then μν=νμ\mu\otimes\nu=\nu\otimes\mu.

Definition 3.8.

[CGH23] Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) and M𝒰M\prec\mathcal{U} a small submodel such that μ\mu is MM-invariant, and 𝐱=(xi)i<ω\mathbf{x}=(x_{i})_{i<\omega}. We say that μ\mu is self-averaging over MM if for any measure λ𝔐𝐱(𝒰)\lambda\in\mathfrak{M}_{\mathbf{x}}(\mathcal{U}) with λ|M=μ(ω)|M\lambda|_{M}=\mu^{(\omega)}|_{M} and any formula φ(x)x(𝒰)\varphi(x)\in\mathcal{L}_{x}(\mathcal{U}) we have

limiλ(φ(xi))=μ(φ(x)).\lim_{i\to\infty}\lambda(\varphi(x_{i}))=\mu(\varphi(x)).

The following generalizes a standard characterization of generically stable measures in NIP theories [HPS13] to fim measures in arbitrary theories (and demonstrates in particular that if pSx(𝒰)p\in S_{x}(\mathcal{U}) is fim  viewed as a Keisler measure, then it is generically stable in the sense of Section 2.1; indeed, pp is generically stable over MM if for every Morley sequence (ai)i<ω(a_{i})_{i<\omega} in pp over MM we have limiωtp(ai/𝒰)=p\lim_{i\to\omega}\operatorname{tp}(a_{i}/\mathcal{U})=p — and if δp\delta_{p} is self-averaging, this property holds, see [CG20, Proposition 3.2]):

Fact 3.9.

[CGH23, Theorem 2.7] If μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) and μ\mu is fim over MM, then μ\mu is self-averaging over MM.

The following would be a natural generalization of stationarity for generically stable types to measures:

Conjecture 3.10.

Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) be fim over M𝒰M\prec\mathcal{U} and let AA be a small set with MA𝒰M\subseteq A\subseteq\mathcal{U}. If ν𝔐x(𝒰)\nu\in\mathfrak{M}_{x}(\mathcal{U}) is AA-invariant, Borel-definable, and ν|A=μ|A\nu|_{A}=\mu|_{A}, then μ=ν\mu=\nu.

Conjecture 3.10 is known to hold when μ=p\mu=p is a type and TT is an arbitrary theory (by Fact 2.1(2)) and when μ\mu is a measure but TT is NIP (by [HPS13, Proposition 3.3]). The following proposition is a special case of Conjecture 3.10 sufficient for our purposes here.

Proposition 3.11.

Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) be fim over MM and AMA\supseteq M. Suppose ν𝔐x(𝒰)\nu\in\mathfrak{M}_{x}(\mathcal{U}) is AA-invariant and ν|A=μ|A\nu|_{A}=\mu|_{A}. If either of the following holds:

  1. (1)

    ν\nu is definable,

  2. (2)

    or the measures μ(n)\mu^{(n)} are fim for each n1n\geq 1 and ν\nu is Borel-definable;

then μ=ν\mu=\nu.

Proof.

The only difference between the two cases is the justification of equation (b)(b) below. Suppose that we are given some ν\nu with the described properties. Since ν\nu is AA-invariant and (Borel-)definable, it follows that ν\nu is (Borel-)definable over AA.

Claim.

We have that ν(ω)|A=μ(ω)|A\nu^{(\omega)}|_{A}=\mu^{(\omega)}|_{A} (in either of the two cases).

Proof.

By assumption ν(1)|A=μ(1)|A\nu^{(1)}|_{A}=\mu^{(1)}|_{A}. Assume that we have already established ν(n)|A=μ(n)|A\nu^{(n)}|_{A}=\mu^{(n)}|_{A} for some nωn\in\omega. Fix an arbitrary formula θ(x1,,xn+1)x1,,xn+1(A)\theta(x_{1},\ldots,x_{n+1})\in\mathcal{L}_{x_{1},\ldots,x_{n+1}}(A) and ε>0\varepsilon>0. Let ρ(xn+1;x1,,xn):=θ(x1,,xn+1)\rho(x_{n+1};x_{1},\ldots,x_{n}):=\theta(x_{1},\ldots,x_{n+1}). Since the measure μ(n)(x1,,xn)\mu^{(n)}(x_{1},\ldots,x_{n}) is definable over AA, there exist formulas ψ1(xn+1,a¯),\psi_{1}(x_{n+1},\bar{a}),\ldots, ψm(xn+1,a¯)xn+1(A)\psi_{m}(x_{n+1},\bar{a})\in\mathcal{L}_{x_{n+1}}(A) and r1,,rm[0,1]r_{1},\ldots,r_{m}\in[0,1] such that

(\dagger) supqSxn+1(A)|Fμ(n),Aρ(q)i=1mri𝟏ψi(xn+1,a¯)(q)|ε.\sup_{q\in S_{x_{n+1}}(A)}\left\lvert F_{\mu^{(n)},A}^{\rho^{*}}(q)-\sum_{i=1}^{m}r_{i}\mathbf{1}_{\psi_{i}(x_{n+1},\bar{a})}(q)\right\rvert\leq\varepsilon.

Then we have:

ν(n+1)(θ(x1,,xn+1))=(νxn+1νx1,,xn(n))(θ(x1,,xn+1))\displaystyle\nu^{(n+1)}(\theta(x_{1},\ldots,x_{n+1}))=\left(\nu_{x_{n+1}}\otimes\nu^{(n)}_{x_{1},\ldots,x_{n}}\right)\left(\theta(x_{1},\ldots,x_{n+1})\right)
=Sx1,,xn(A)Fνxn+1,Aρd(ν(n)|A)=(a)Sx1,,xn(A)Fνxn+1,Aρd(μ(n)|A)\displaystyle=\int_{S_{x_{1},\ldots,x_{n}}(A)}F_{\nu_{x_{n+1}},A}^{\rho}\ \mathrm{d}\left(\nu^{(n)}|_{A}\right)\overset{(a)}{=}\int_{S_{x_{1},\ldots,x_{n}}(A)}F_{\nu_{x_{n+1}},A}^{\rho}\ \mathrm{d}\left(\mu^{(n)}|_{A}\right)
=(b)Sxn+1(A)Fμ(n),Aρd(ν|A)(c)εSxn+1(A)i=1mri𝟏ψi(xn+1,a¯)d(ν|A)\displaystyle\overset{(b)}{=}\int_{S_{x_{n+1}}(A)}F_{\mu^{(n)},A}^{\rho^{*}}\ \textrm{d}\left(\nu|_{A}\right)\overset{(c)}{\approx}_{\varepsilon}\int_{S_{x_{n+1}}(A)}\sum_{i=1}^{m}r_{i}\mathbf{1}_{\psi_{i}(x_{n+1},\bar{a})}\ \textrm{d}\left(\nu|_{A}\right)
=i=1mriν(ψi(xn+1,a¯))=(d)i=1mriμ(ψi(xn+1,a¯))\displaystyle=\sum_{i=1}^{m}r_{i}\nu(\psi_{i}(x_{n+1},\bar{a}))\overset{(d)}{=}\sum_{i=1}^{m}r_{i}\mu(\psi_{i}(x_{n+1},\bar{a}))
=Sxn+1(A)i=1mri𝟏ψi(xn+1,a¯)d(μ|A)(c)εSxn+1(A)Fμ(n),Aρd(μ|A)\displaystyle=\int_{S_{x_{n+1}}(A)}\sum_{i=1}^{m}r_{i}\mathbf{1}_{\psi_{i}(x_{n+1},\bar{a})}\ \textrm{d}\left(\mu|_{A}\right)\overset{(c)}{\approx}_{\varepsilon}\int_{S_{x_{n+1}}(A)}F_{\mu^{(n)},A}^{\rho^{*}}\ \textrm{d}\left(\mu|_{A}\right)
=(μx1,,xn(n)μxn+1)(θ(x1,,xn+1))=(e)(μxn+1μx1,,xn(n))(θ(x1,,xn+1))\displaystyle=\left(\mu^{(n)}_{x_{1},\ldots,x_{n}}\otimes\mu_{x_{n+1}}\right)(\theta(x_{1},\ldots,x_{n+1}))\overset{(e)}{=}\left(\mu_{x_{n+1}}\otimes\mu^{(n)}_{x_{1},\ldots,x_{n}}\right)(\theta(x_{1},\ldots,x_{n+1}))
=μ(n+1)(θ(x1,,xn+1)),\displaystyle=\mu^{(n+1)}(\theta(x_{1},\ldots,x_{n+1})),

with the following justifications for the corresponding steps:

  1. (a)

    induction hypothesis;

  2. (b)

    in Case (1), μ(n)\mu^{(n)} is fam over AA since μ(n)\mu^{(n)} is fam over MM and MAM\subseteq A (and fim  implies fam over a model), by [CG20, Proposition 2.10(b)], ν\nu is definable over AA, and fam measures commute with definable measures [CGH23a, Proposition 5.17]; in Case (2), μ(n)\mu^{(n)} is fim over AA, ν\nu is Borel-definable over AA, and fim measures commute with Borel-definable measures [CGH23a, Proposition 5.15];

  3. (c)

    by ()(\dagger);

  4. (d)

    by assumption;

  5. (e)

    μxn+1\mu_{x_{n+1}} is fim, and fim measures commute with all Borel-definable measures [CGH23a, Proposition 5.15] (alternatively, fam measures commute with definable measures).

As θ\theta and ε\varepsilon were arbitrary, we conclude ν(n+1)|A=μ(n+1)|A\nu^{(n+1)}|A=\mu^{(n+1)}|_{A}. And then ν(ω)|A=n<ων(n)=n<ωμ(n)=μ(ω)|A\nu^{(\omega)}|_{A}=\bigcup_{n<\omega}\nu^{(n)}=\bigcup_{n<\omega}\mu^{(n)}=\mu^{(\omega)}|_{A}. ∎

Let now φ(x)x(𝒰)\varphi(x)\in\mathcal{L}_{x}(\mathcal{U}) be arbitrary. Let λ(𝐱):=ν(ω)(𝐱)\lambda(\mathbf{x}):=\nu^{(\omega)}(\mathbf{x}), by the claim above we have λ|A=μ(ω)|A\lambda|_{A}=\mu^{(\omega)}|_{A}. Note that for every iωi\in\omega, λ(φ(xi))=ν(φ(x))\lambda(\varphi(x_{i}))=\nu(\varphi(x)). As μ\mu is fim over MM, it is self-averaging over MM by Fact 3.9, hence ν(φ(x))=limiλ(φ(xi))=μ(φ(x))\nu(\varphi(x))=\lim_{i\to\infty}\lambda(\varphi(x_{i}))=\mu(\varphi(x)). ∎

Remark 3.12.

Our proof of Proposition 3.11 does not apply to the general case of Conjecture 3.10 since it is open whether or not μ\mu being fim implies that μ(n)\mu^{(n)} is fim for n2n\geq 2 in an arbitrary theory, even when μ=p\mu=p is a generically stable type.

3.4. Fim measures over “random” parameters

In this section we prove a generalization of Fact 3.9 of independent interest, demonstrating that any Morley sequence of a fim  measure determines the measure of arbitrary formulas by averaging along it — even when the parameters of these formulas are allowed to be “random”. More precisely:

Theorem 3.13.

Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) be fim over MM, ν𝔐y(𝒰)\nu\in\mathfrak{M}_{y}(\mathcal{U}), φ(x,y,z)xyz\varphi(x,y,z)\in\mathcal{L}_{xyz}, b𝒰zb\in\mathcal{U}^{z}, and 𝐱=(xi)iω\mathbf{x}=(x_{i})_{i\in\omega}. Suppose that λ𝔐𝐱y(𝒰)\lambda\in\mathfrak{M}_{\mathbf{x}y}(\mathcal{U}) is arbitrary such that λ|𝐱,M=μ(ω)\lambda|_{\mathbf{x},M}=\mu^{(\omega)} and λ|y=ν\lambda|_{y}=\nu. Then

limiλ(φ(xi,y,b))=μν(φ(x,y,b)).\lim_{i\to\infty}\lambda(\varphi(x_{i},y,b))=\mu\otimes\nu(\varphi(x,y,b)).

Moreover for every ε>0\varepsilon>0 there exists n=n(μ,φ,ε)n=n(\mu,\varphi,\varepsilon)\in\mathbb{N} so that for any ν,λ,b\nu,\lambda,b as above, we have λ(φ(xi,y,b))εμν(φ(x,y,b))\lambda(\varphi(x_{i},y,b))\approx^{\varepsilon}\mu\otimes\nu(\varphi(x,y,b)) for all but nn many ii\in\mathbb{N}.

Remark 3.14.

Fact 3.9 corresponds to the special case when ν(y)=q(y)\nu(y)=q(y) is a type. We note that this result is central to the proof of our main theorem, and is new even for NIP theories.

Our proof of Theorem 3.13 relies on the use of Keisler randomization in continuous logic, as introduced and studied in [BYK09]. We will follow the notation from [CGH23, Section 3.2]. Let TT be a complete first order theory. We let TRT^{R} denote the (continuous) first order theory of its Keisler randomization (we refer to [BYK09, Section 2] for the details). Let MM be a model of TT and let (Ω,,)(\Omega,\mathcal{B},\mathbb{P}) be a probability algebra. We consider the model M(Ω,,)M^{(\Omega,\mathcal{B},\mathbb{P})} of TRT^{R}, which we usually denote as MΩM^{\Omega} for brevity, defined as follows. We let

M0:={f:ΩM:f is -measurable,|im(f)|<0},M^{\prime}_{0}:=\left\{f:\Omega\to M:f\text{ is }\mathcal{B}\text{-measurable},|\operatorname{im}(f)|<\aleph_{0}\right\},

equipped with the pseudo-metric d(f,g):=({ωΩ:f(ω)g(ω)})d(f,g):=\mathbb{P}\left(\{\omega\in\Omega:f(\omega)\neq g(\omega)\}\right). Then MΩM^{\Omega} is constructed by taking the metric completion of M0M^{\prime}_{0} and then identifying random variables up to \mathbb{P}-measure 0. We let M0ΩM^{\Omega}_{0} be the set of classes of elements of M0M_{0}^{\prime}. By construction, M0ΩM^{\Omega}_{0} is a metrically dense (pre-)substructure of MΩM^{\Omega}.

Let 𝒰\mathcal{U} be a monster model of TT such that M𝒰M\prec\mathcal{U}. The model 𝒰Ω\mathcal{U}^{\Omega} is almost never saturated, so we will always think of 𝒰Ω\mathcal{U}^{\Omega} as (elementarily) embedded into a monster model 𝒞\mathcal{C} of TRT^{R}, i.e. 𝒰Ω𝒞\mathcal{U}^{\Omega}\prec\mathcal{C}. If a𝒰a\in\mathcal{U}, we let fa𝒰0Ωf_{a}\in\mathcal{U}^{\Omega}_{0} denote the constant random variable taking value aa, i.e. faf_{a} is the equivalence class of the maps which send Ω\Omega to the point aa (equivalence up to measure 0). If A𝒰A\subseteq\mathcal{U}, we let Ac:={fa:aA}𝒰0ΩA^{c}:=\left\{f_{a}:a\in A\right\}\subseteq\mathcal{U}^{\Omega}_{0}. If φ(x1,,xn)\varphi(x_{1},\ldots,x_{n}) is an \mathcal{L}-formula, we let 𝔼[φ(x1,,xn)]\mathbb{E}[\varphi(x_{1},\ldots,x_{n})] denote the corresponding continuous formula in the randomization. This formula is evaluated on tuples of elements h¯=(h1,,hn)\bar{h}=(h_{1},\ldots,h_{n}) from 𝒰0Ω\mathcal{U}^{\Omega}_{0} via

𝔼[φ(h¯)]=({ωΩ:𝒰φ(h1(ω),,hn(ω))}),\mathbb{E}\left[\varphi(\bar{h})\right]=\mathbb{P}\left(\{\omega\in\Omega:\mathcal{U}\models\varphi(h_{1}(\omega),\ldots,h_{n}(\omega))\}\right),

and is extended to 𝒰Ω\mathcal{U}^{\Omega} via uniform limits. For B𝒞B\subseteq\mathcal{C}, SxR(B)S^{R}_{x}(B) will denote the space of types in the tuple of variables xx over BB in TRT^{R}.

Remark 3.15.

Note that for any h¯=(h1,,hn)(𝒰0Ω)n\bar{h}=(h_{1},\ldots,h_{n})\in(\mathcal{U}^{\Omega}_{0})^{n}, there exists a finite \mathcal{B}-measurable partition 𝒜\mathcal{A} of Ω\Omega with the property that for each ini\leq n the function hih_{i} is constant on each element of 𝒜\mathcal{A}. Given such an h¯\bar{h} and 𝒜\mathcal{A}, we write h¯|A\bar{h}|_{A} for the tuple of constant values of the functions in h¯\bar{h} on the set AA. Note that for each A𝒜A\in\mathcal{A}, h¯|A\bar{h}|_{A} is an element of 𝒰n\mathcal{U}^{n}.

The following fact can be derived from basic facts about continuous logic:

Fact 3.16.

Suppose that pSxR(𝒰Ω)p\in S_{x}^{R}(\mathcal{U}^{\Omega}). Then there exists a net of tuples (hi)iI(h_{i})_{i\in I} where hi(𝒰0Ω)xh_{i}\in(\mathcal{U}^{\Omega}_{0})^{x} such that limiItpR(hi/𝒰Ω)=p\lim_{i\in I}\operatorname{tp}^{R}(h_{i}/\mathcal{U}^{\Omega})=p.

The following observations were made by Ben Yaacov in an unpublished note [BY09]. For a detailed verification, we refer the reader to [CGH23, Section 3.2].

Fact 3.17.

Let 𝒰\mathcal{U} be a monster model of TT, μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}), and 𝒰Ω𝒞\mathcal{U}^{\Omega}\prec\mathcal{C}.

  1. (1)

    There exists a unique type pμSxR(𝒰Ω)p_{\mu}\in S_{x}^{R}(\mathcal{U}^{\Omega}) such that for any \mathcal{L}-formula φ(x,y¯)\varphi(x,\bar{y}), h¯=(h1,,hn)𝒰0Ω\bar{h}=(h_{1},\ldots,h_{n})\in\mathcal{U}^{\Omega}_{0}, and any measurable partition 𝒜\mathcal{A} such that each element of h¯\bar{h} is constant on each element of 𝒜\mathcal{A},

    (𝔼[φ(x,h¯)])pμ=A𝒜(A)μ(φ(x,h¯|A)).(\mathbb{E}[\varphi(x,\bar{h})])^{p_{\mu}}=\sum_{A\in\mathcal{A}}\mathbb{P}(A)\mu(\varphi(x,\bar{h}|_{A})).
  2. (2)

    If μ\mu is definable, then there exists a unique type rμSxR(𝒞)r_{\mu}\in S_{x}^{R}(\mathcal{C}) such that:

    1. (a)

      rμ|𝒰Ω=pμr_{\mu}|_{\mathcal{U}^{\Omega}}=p_{\mu};

    2. (b)

      rμr_{\mu} is definable over 𝒰Ω{\mathcal{U}^{\Omega}}; if μ\mu is MM-definable, then rμr_{\mu} is MΩM^{\Omega}-definable.

Remark 3.18.

The claims in Fact 3.17 hold in the context where xx is an infinite tuple of variables. The infinitary results follow easily from their finite counterparts.

Corollary 3.19.

Suppose that 𝐱=(xi)i<α\mathbf{x}=(x_{i})_{i<\alpha}, 𝐲=(yi)i<β\mathbf{y}=(y_{i})_{i<\beta}, and λ𝔐𝐱𝐲(𝒰)\lambda\in\mathfrak{M}_{\mathbf{xy}}(\mathcal{U}). Then

(pλ)|𝐱,MΩ=p(λ|𝐱)|MΩ and (pλ)|𝐲=p(λ|𝐲).\left(p_{\lambda}\right)|_{\mathbf{x},M^{\Omega}}=p_{\left(\lambda|_{\mathbf{x}}\right)}|_{M^{\Omega}}\textrm{ and }\left(p_{\lambda}\right)|_{\mathbf{y}}=p_{\left(\lambda|_{\mathbf{y}}\right)}.
Proof.

Fix x¯=(xi1,,xin)\bar{x}=\left(x_{i_{1}},\ldots,x_{i_{n}}\right) and h¯:=(h1,,hm)(MΩ)m\overline{h}:=\left(h_{1},\ldots,h_{m}\right)\in\left(M^{\Omega}\right)^{m}. Fix a finite measurable partition 𝒜\mathcal{A} of Ω\Omega such that each element of h¯\overline{h} is constant on each element of 𝒜\mathcal{A}. From the definitions we have:

(𝔼[φ(x¯,h¯)])pλ\displaystyle\left(\mathbb{E}\left[\varphi\left(\bar{x},\overline{h}\right)\right]\right)^{p_{\lambda}} =A𝒜(A)λ(φ(x¯,h¯|A))\displaystyle=\sum_{A\in\mathcal{A}}\mathbb{P}(A)\lambda(\varphi(\bar{x},\overline{h}|_{A}))
=A𝒜(A)(λ|𝐱)(φ(x¯,h¯|A))\displaystyle=\sum_{A\in\mathcal{A}}\mathbb{P}(A)\left(\lambda|_{\mathbf{x}}\right)(\varphi(\bar{x},\overline{h}|_{A}))
=(𝔼[φ(x¯,h¯)])p(λ|𝐱).\displaystyle=(\mathbb{E}[\varphi(\bar{x},\overline{h})])^{p_{\left(\lambda|_{\mathbf{x}}\right)}}.

By quantifier elimination in TRT^{R}, we conclude that pλ|𝐱,MΩ=p(λ𝐱)|MΩp_{\lambda}|_{\mathbf{x},M^{\Omega}}=p_{\left(\lambda_{\mathbf{x}}\right)}|_{M^{\Omega}}. ∎

We recall some results from [CGH23] connecting the randomized measures, the Morley product and generic stability in TT and TRT^{R}. These are [CGH23, Proposition 3.15], [CGH23, Corollary 3.16] and [CGH23, Corollary 3.19], respectively.

Fact 3.20.

Suppose μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) and ν𝔐y(𝒰)\nu\in\mathfrak{M}_{y}(\mathcal{U}).

  1. (1)

    If μ\mu and ν\nu are definable, then

    rμν(x,y)=rμ(x)rν(y).r_{\mu\otimes\nu}(x,y)=r_{\mu}(x)\otimes r_{\nu}(y).
  2. (2)

    If μ\mu is definable, then for every n1n\geq 1,

    rμ(n)(x¯)=(rμ)(n)(x¯).r_{\mu^{(n)}}(\bar{x})=(r_{\mu})^{(n)}(\bar{x}).
  3. (3)

    If μ\mu is fim, then rμr_{\mu} is generically stable over MΩM^{\Omega} (for generically stable types in continuous logic we refer to [Kha22, CGH23, And23]).

Corollary 3.21.

If μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) is a definable measure then rμ(ω)=(rμ)(ω)r_{\mu^{(\omega)}}=(r_{\mu})^{(\omega)}.

Proof.

First note

(rμ)(ω)=1n<ωrμ(n)=1n<ωrμ(n).(r_{\mu})^{(\omega)}=\bigcup_{1\leq n<\omega}r_{\mu}^{(n)}=\bigcup_{1\leq n<\omega}r_{\mu^{(n)}}.

We want to show that n<ωrμ(n)=rμ(ω)\bigcup_{n<\omega}r_{\mu^{(n)}}=r_{\mu^{(\omega)}}. By quantifier elimination in TRT^{R}, it suffices to show that for every x¯,y\mathcal{L}_{\bar{x},y}-formula φ(x1,,xk,y)\varphi(x_{1},\ldots,x_{k},y) and b𝒞yb\in\mathcal{C}^{y}, we have that 𝔼[φ(x¯,b)]rμ(ω)=𝔼[φ(x¯,b)]rμ(k)\mathbb{E}[\varphi(\bar{x},b)]^{r_{\mu^{(\omega)}}}=\mathbb{E}[\varphi(\bar{x},b)]^{r_{\mu^{(k)}}}. By Fact 3.16, fix a net (hi)iI(h_{i})_{i\in I} of elements each in (𝒰0Ω)y(\mathcal{U}^{\Omega}_{0})^{y} such that limiItpR(hi/𝒰Ω)=tpR(b/𝒰Ω)\lim_{i\in I}\operatorname{tp}^{R}(h_{i}/\mathcal{U}^{\Omega})=\operatorname{tp}^{R}(b/\mathcal{U}^{\Omega}). For each iIi\in I, choose a finite measurable partition 𝒜i\mathcal{A}_{i} of Ω\Omega such that each element of hih_{i} is constant on each element of 𝒜i\mathcal{A}_{i}. We have the following computation (using Fact 3.17):

𝔼[φ(x¯,b)]rμ(ω)=limiIFrμ(ω)𝔼[φ(x¯,y)]((tpR(hi/𝒰Ω))\displaystyle\mathbb{E}[\varphi(\bar{x},b)]^{r_{\mu^{(\omega)}}}=\lim_{i\in I}F_{r_{\mu^{(\omega)}}}^{\mathbb{E}[\varphi(\bar{x},y)]}\left((\operatorname{tp}^{R}(h_{i}/\mathcal{U}^{\Omega})\right)
=limiI𝔼[φ(x¯,hi)]pμ(ω)=limiIA𝒜i(A)μ(ω)(φ(x¯,hi|A))\displaystyle=\lim_{i\in I}\mathbb{E}[\varphi(\bar{x},h_{i})]^{p_{\mu^{(\omega)}}}=\lim_{i\in I}\sum_{A\in\mathcal{A}_{i}}\mathbb{P}(A)\mu^{(\omega)}(\varphi(\bar{x},h_{i}|_{A}))
=limiIA𝒜i(A)μ(k)(φ(x¯,hi|A))=limiI(𝔼[φ(x¯,hi)])pμ(k)\displaystyle=\lim_{i\in I}\sum_{A\in\mathcal{A}_{i}}\mathbb{P}(A)\mu^{(k)}(\varphi(\bar{x},h_{i}|_{A}))=\lim_{i\in I}(\mathbb{E}[\varphi(\bar{x},h_{i})])^{p_{\mu^{(k)}}}
=limiIFrμ(k)𝔼[φ(x¯,y)]((tpR(hi/𝒰Ω))=𝔼[φ(x¯,b)]rμ(k),\displaystyle=\lim_{i\in I}F_{r_{\mu^{(k)}}}^{\mathbb{E}[\varphi(\bar{x},y)]}\left((\operatorname{tp}^{R}(h_{i}/\mathcal{U}^{\Omega})\right)=\mathbb{E}[\varphi(\bar{x},b)]^{r_{\mu^{(k)}}},

where the first and last equality follow from the fact that Frμ(ω)𝔼[φ(x¯,y)]F_{r_{\mu^{(\omega)}}}^{\mathbb{E}[\varphi(\bar{x},y)]} and Frμ(k)𝔼[φ(x¯,y)]F_{r_{\mu^{(k)}}}^{\mathbb{E}[\varphi(\bar{x},y)]} are continuous maps, by definability of μ\mu, hence of μ(ω)\mu^{(\omega)} and μ(k)\mu^{(k)}, and Fact 3.17(2)(b). ∎

The following fact is [CGH23, Lemma 3.13].

Fact 3.22.

Suppose that μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) and (hi)iI(h_{i})_{i\in I} is a net of elements such that hi(𝒰0Ω)xh_{i}\in(\mathcal{U}^{\Omega}_{0})^{x} and limiItpR(hi/𝒰Ω)=pμ\lim_{i\in I}\operatorname{tp}^{R}(h_{i}/\mathcal{U}^{\Omega})=p_{\mu}. For each iIi\in I, let 𝒜i\mathcal{A}_{i} be a finite measurable partition of Ω\Omega such that each element of hih_{i} is constant on each AAiA\in A_{i}. Then

limiI(A𝒜i(A)δ(hi|A))=μ,\lim_{i\in I}\left(\sum_{A\in\mathcal{A}_{i}}\mathbb{P}(A)\delta_{(h_{i}|_{A})}\right)=\mu,

where the limit is calculated in the space 𝔐x(𝒰)\mathfrak{M}_{x}(\mathcal{U}).

Proposition 3.23.

Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) be fim over a small model M𝒰M\prec\mathcal{U}, ν𝔐y(𝒰)\nu\in\mathfrak{M}_{y}(\mathcal{U}) and 𝐱=(xi)iω\mathbf{x}=(x_{i})_{i\in\omega} with xix_{i} of the same sort as xx for all ii. Suppose that λ𝔐𝐱y(𝒰)\lambda\in\mathfrak{M}_{\mathbf{x}y}(\mathcal{U}) such that λ|𝐱,M=μ(ω)|M\lambda|_{\mathbf{x},M}=\mu^{(\omega)}|_{M} and λ|y=ν\lambda|_{y}=\nu. Then for any φ(x,y)\varphi(x,y)\in\mathcal{L} we have

limiλ(φ(xi,y))=μν(φ(x,y)).\lim_{i\to\infty}\lambda(\varphi(x_{i},y))=\mu\otimes\nu(\varphi(x,y)).
Proof.

The measures μ𝔐x(𝒰),λ𝔐𝐱y(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}),\lambda\in\mathfrak{M}_{\mathbf{x}y}(\mathcal{U}), μ(ω)𝔐𝐱(𝒰)\mu^{(\omega)}\in\mathfrak{M}_{\mathbf{x}}(\mathcal{U}) and ν𝔐y(𝒰)\nu\in\mathfrak{M}_{y}(\mathcal{U}) can be associated to complete types pμSxR(𝒰Ω),pλS𝐱yR(𝒰Ω)p_{\mu}\in S^{R}_{x}(\mathcal{U}^{\Omega}),p_{\lambda}\in S^{R}_{\mathbf{x}y}(\mathcal{U}^{\Omega}), pμ(ω)S𝐱yR(𝒰Ω)p_{\mu^{(\omega)}}\in S^{R}_{\mathbf{x}y}(\mathcal{U}^{\Omega}) and pνSyR(𝒰Ω)p_{\nu}\in S^{R}_{y}(\mathcal{U}^{\Omega}), respectively, by Fact 3.17(1). As μ\mu is MM-definable, μ(ω)\mu^{(\omega)} is also MM-definable, and so the types rμSx(𝒞)r_{\mu}\in S_{x}(\mathcal{C}) and rμ(ω)S𝐱(𝒞)r_{\mu^{(\omega)}}\in S_{\mathbf{x}}(\mathcal{C}) are well-defined and definable over MΩM^{\Omega} by Fact 3.17(2). We then have:

  1. (1)

    rμr_{\mu} is generically stable over MΩM^{\Omega} (by Fact 3.20(3));

  2. (2)

    rμ(ω)|MΩ=(rμ)(ω)|MΩr_{\mu^{(\omega)}}|_{M^{\Omega}}=(r_{\mu})^{(\omega)}|_{M^{\Omega}} (by Corollary 3.21);

  3. (3)

    pλ|𝐱,MΩ=pμ(ω)|MΩ=rμ(ω)|MΩp_{\lambda}|_{\mathbf{x},M^{\Omega}}=p_{\mu^{(\omega)}}|_{M^{\Omega}}=r_{\mu^{(\omega)}}|_{M^{\Omega}} and pλ|y=pνp_{\lambda}|_{y}=p_{\nu} (by Corollary 3.19).

Let 𝐚=(ai)iω\mathbf{a}=(a_{i})_{i\in\omega} and bb in 𝒞\mathcal{C} be so that (𝐚,b)pλ(\mathbf{a},b)\models p_{\lambda}. Then 𝐚\mathbf{a} is a Morley sequence in rμr_{\mu} over MΩM^{\Omega} by (2) and (3) from above. By Fact 3.16, choose a net (hj)jJ(h_{j})_{j\in J} of tuples in (𝒰0)y(\mathcal{U}_{0})^{y} such that limjJtpR(hj/𝒰Ω)=tpR(b/𝒰Ω)=pν\lim_{j\in J}\operatorname{tp}^{R}(h_{j}/\mathcal{U}^{\Omega})=\operatorname{tp}^{R}(b/\mathcal{U}^{\Omega})=p_{\nu}. For each hjh_{j}, choose a finite measurable partition 𝒜j\mathcal{A}_{j} of Ω\Omega such that each function in hjh_{j} is constant on each element of 𝒜j\mathcal{A}_{j}. Now, given any φ(x,y)\varphi(x,y)\in\mathcal{L}, we then have the following computation:

limiλ(φ(xi,y))=limi𝔼[φ(xi,y)]pλ=limi𝔼[φ(ai,b)]=(a)𝔼[φ(x,b)]rμ\displaystyle\lim_{i\to\infty}\lambda(\varphi(x_{i},y))=\lim_{i\to\infty}\mathbb{E}[\varphi(x_{i},y)]^{p_{\lambda}}=\lim_{i\to\infty}\mathbb{E}[\varphi(a_{i},b)]\overset{(a)}{=}\mathbb{E}[\varphi(x,b)]^{r_{\mu}}
=Frμφ(tpR(b/𝒰Ω))=(b)limjJFrμφ(tpR(hj/𝒰Ω))=limjJ𝔼[φ(x,hj)]pμ\displaystyle=F^{\varphi}_{r_{\mu}}\left(\operatorname{tp}^{R}(b/\mathcal{U}^{\Omega})\right)\overset{(b)}{=}\lim_{j\in J}F^{\varphi}_{r_{\mu}}\left(\operatorname{tp}^{R}(h_{j}/\mathcal{U}^{\Omega})\right)=\lim_{j\in J}\mathbb{E}[\varphi(x,h_{j})]^{p_{\mu}}
=(c)limjJA𝒜j(A)μ(φ(x,hj|A))=(d)limjJSy(𝒰)Fμφd(A𝒜j(A)δ(hj|A))\displaystyle\overset{(c)}{=}\lim_{j\in J}\sum_{A\in\mathcal{A}_{j}}\mathbb{P}(A)\mu(\varphi(x,h_{j}|_{A}))\overset{(d)}{=}\lim_{j\in J}\int_{S_{y}(\mathcal{U})}F_{\mu}^{\varphi}\ \textrm{d}\left(\sum_{A\in\mathcal{A}_{j}}\mathbb{P}(A)\delta_{\left(h_{j}|_{A}\right)}\right)
=(e)Sy(𝒰)Fμφd(limjJA𝒜j(A)δ(hj|A))=(f)Sy(𝒰)Fμφ𝑑ν=μν(φ(x,y)),\displaystyle\overset{(e)}{=}\int_{S_{y}(\mathcal{U})}F_{\mu}^{\varphi}\ \textrm{d}\left(\lim_{j\in J}\sum_{A\in\mathcal{A}_{j}}\mathbb{P}(A)\delta_{\left(h_{j}|_{A}\right)}\right)\overset{(f)}{=}\int_{S_{y}(\mathcal{U})}F_{\mu}^{\varphi}d\nu=\mu\otimes\nu(\varphi(x,y)),

where the corresponding equalities hold for the following reasons:

  1. (a)(a)

    since the type rμSxR(𝒞)r_{\mu}\in S_{x}^{R}(\mathcal{C}) is generically stable over MΩM^{\Omega} by (1) and (𝐚i)iω(\mathbf{a}_{i})_{i\in\omega} is a Morley sequence in rμr_{\mu} over MΩM^{\Omega};

  2. (b)(b)

    by the choice of (hj)jI(h_{j})_{j\in I} and, as rμr_{\mu} is definable over 𝒰Ω\mathcal{U}^{\Omega} by (1), the map Frμφ:SyR(𝒰Ω)[0,1]F_{r_{\mu}}^{\varphi}:S^{R}_{y}(\mathcal{U}^{\Omega})\to[0,1] is continuous;

  3. (c)(c)

    by the definition of pμp_{\mu} (Fact 3.17);

  4. (d)(d)

    for a fixed jJj\in J, the computations of the left hand side and the right hand side are the same;

  5. (e)(e)

    since μ\mu is definable, the map Fμφ:Sy(𝒰)[0,1]F_{\mu}^{\varphi}:S_{y}(\mathcal{U})\to[0,1] is continuous, hence the map γ𝔐y(𝒰)Fμφdγ[0,1]\gamma\in\mathfrak{M}_{y}(\mathcal{U})\mapsto\int F_{\mu}^{\varphi}\textrm{d}\gamma\in[0,1] is continuous;

  6. (f)(f)

    by Fact 3.22. ∎

We will use the following general topological fact [CGH23, Lemma 2.3]:

Fact 3.24.

Let f:XKf:X\to K be an arbitrary function from a compact Hausdorff space to a compact interval KK\subseteq\mathbb{R}. Suppose there is a closed subset CKω×XC\subseteq K^{\omega}\times X satisfying the following properties:

  1. (1)

    the projection of CC onto XX is all of XX;

  2. (2)

    if (α,x)C(\alpha,x)\in C and g:ωωg:\omega\to\omega is strictly increasing, then (αg,x)C(\alpha\circ g,x)\in C;

  3. (3)

    for any (α,x)C(\alpha,x)\in C, limiα(i)=f(x)\lim_{i\to\infty}\alpha(i)=f(x).

Then ff is continuous and, for any ε>0\varepsilon>0, there is an nεn_{\varepsilon}\in\mathbb{N} such that: for any (α,x)C(\alpha,x)\in C, {iω:α(i)εf(x)}nε\{i\in\omega:\alpha(i)\not\approx_{\varepsilon}f(x)\}\leq n_{\varepsilon}.

Finally, we can derive the main theorem of the section:

Proof of Theorem 3.13.

The “moreover” clause of Theorem 3.13 for formulas without parameters follows from Proposition 3.23 by compactness (using Fact 3.24).

Namely, first let φ(x;y,z)()\varphi(x;y,z)\in\mathcal{L}(\emptyset) be arbitrary. We let

𝔐L(𝒰):={η𝔐𝐱yz(𝒰):η|𝐱,M=μ(ω)|M},X:=𝔐yz(𝒰),K=[0,1],\displaystyle\mathfrak{M}_{L}(\mathcal{U}):=\left\{\eta\in\mathfrak{M}_{\mathbf{x}yz}(\mathcal{U}):\eta|_{\mathbf{x},M}=\mu^{(\omega)}|_{M}\right\},X:=\mathfrak{M}_{yz}\left(\mathcal{U}\right),K=[0,1],
f:XK defined by ν𝔐yz(𝒰)μν(φ(x,y,z)), and\displaystyle f:X\to K\textrm{ defined by }\nu\in\mathfrak{M}_{yz}\left(\mathcal{U}\right)\mapsto\mu\otimes\nu(\varphi(x,y,z)),\textrm{ and}
C:={((η(φ(xi,y,z)):iω),η|yz):η𝔐L(𝒰)}[0,1]ω×X.\displaystyle C:=\left\{\left(\left(\eta(\varphi(x_{i},y,z)):i\in\omega\right),\eta|_{yz}\right):\eta\in\mathfrak{M}_{L}(\mathcal{U})\right\}\subseteq[0,1]^{\omega}\times X.

The assumptions of Fact 3.24 are satisfied. Indeed, (1) holds since for every ν𝔐yz(𝒰)\nu\in\mathfrak{M}_{yz}\left(\mathcal{U}\right), η:=μ(ω)ν\eta:=\mu^{(\omega)}\otimes\nu gives an element in CC projecting onto it. (2) For every strictly increasing g:ωωg:\omega\to\omega we have a continuous map g:S𝐱(𝒰)S𝐱(𝒰)g^{\prime}:S_{\mathbf{x}}(\mathcal{U})\to S_{\mathbf{x}}(\mathcal{U}) defined by φ(x1,,xn)g(p)φ(xg(1),,xg(n))p\varphi(x_{1},\ldots,x_{n})\in g^{\prime}(p)\iff\varphi(x_{g(1)},\ldots,x_{g(n)})\in p. Now if η|𝐱,M=μ(ω)|M\eta|_{\mathbf{x},M}=\mu^{(\omega)}|_{M} then still g(η)|𝐱,M=μ(ω)|Mg_{\ast}(\eta)|_{\mathbf{x},M}=\mu^{(\omega)}|_{M}, where g(η)g_{\ast}(\eta) is the pushforward of the measure η\eta by gg (see Definition 3.25). And (3) holds by Proposition 3.23. Then we obtain the required n=n(μ,φ,ε)n=n(\mu,\varphi,\varepsilon) applying Fact 3.24.

Now assume we are given μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}), ε>0\varepsilon>0, and ψ(x,y)(𝒰)\psi(x,y)\in\mathcal{L}(\mathcal{U}) is an arbitrary formula with parameters, say of the form φ(x,y,b)\varphi(x,y,b) for some b𝒰zb\in\mathcal{U}^{z} and φ(x,y,z)()\varphi(x,y,z)\in\mathcal{L}(\emptyset). Let n=n(μ,φ,ε)ωn=n(\mu,\varphi,\varepsilon)\in\omega be as given by the above for the formula φ(x,y,z)\varphi(x,y,z) without parameters. Given any λ𝔐𝐱y(𝒰)\lambda\in\mathfrak{M}_{\mathbf{x}y}(\mathcal{U}) with λ|𝐱,M=μ(ω)\lambda|_{\mathbf{x},M}=\mu^{(\omega)} and λ|y=ν\lambda|_{y}=\nu, consider the measures λb𝔐𝐱yz(𝒰)\lambda_{b}\in\mathfrak{M}_{\mathbf{x}yz}(\mathcal{U}) defined by λb(𝐱,y,z):=λ(𝐱,y)δb(z)\lambda_{b}(\mathbf{x},y,z):=\lambda(\mathbf{x},y)\otimes\delta_{b}(z), and νb𝔐yz(𝒰)\nu_{b}\in\mathfrak{M}_{yz}(\mathcal{U}) defined by νb(y,z)=ν(y)δb(z)\nu_{b}(y,z)=\nu(y)\otimes\delta_{b}(z). Note that λb|yz=νb\lambda_{b}|_{yz}=\nu_{b}. By the choice of nn and the previous paragraph (with λb\lambda_{b} and νb\nu_{b} in place of λ\lambda and ν\nu) we have

limiλ(φ(xi,y,b))=limiλb(φ(xi,y,z))=μνb(φ(x,y,z))=μν(φ(x,y,b)).\lim_{i\to\infty}\lambda(\varphi(x_{i},y,b))=\lim_{i\to\infty}\lambda_{b}(\varphi(x_{i},y,z))=\mu\otimes\nu_{b}(\varphi(x,y,z))=\mu\otimes\nu(\varphi(x,y,b)).\qed

3.5. Definable pushforwards of Keisler measures

We record some basic facts about definable pushforwards of Keisler measures.

Definition 3.25.

Let f:𝒰x𝒰yf:\mathcal{U}^{x}\to\mathcal{U}^{y} be a definable map. For μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}), we define the push-forward measure f(μ)f_{*}(\mu) in 𝔐y(𝒰)\mathfrak{M}_{y}(\mathcal{U}), where for any formula φ(y)y(𝒰),f(μ)(φ(y))=μ(φ(f(x)))\varphi(y)\in\mathcal{L}_{y}(\mathcal{U}),f_{*}(\mu)(\varphi(y))=\mu(\varphi(f(x))).

Proposition 3.26.

Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}). Let A𝒰A\subseteq\mathcal{U} be a small set and let f:𝒰x𝒰yf:\mathcal{U}^{x}\to\mathcal{U}^{y} be an AA-definable map. Then we have the following:

  1. (1)

    if μ\mu is AA-invariant, then f(μ)f_{\ast}(\mu) is AA-invariant;

  2. (2)

    if μ\mu is AA-definable, then f(μ)f_{\ast}(\mu) is AA-definable;

  3. (3)

    if μ\mu is fim over AA, then f(μ)f_{\ast}(\mu) is fim over AA;

  4. (4)

    {f(p):pS(μ)}S(f(μ))\left\{f_{\ast}(p):p\in S(\mu)\right\}\subseteq S\left(f_{\ast}(\mu)\right), and if ff is a bijection then these sets are equal.

Proof.

(1) Straightforward.

(2) Note that for any formula φ(y,z)yz\varphi(y,z)\in\mathcal{L}_{yz}, we have Ff(μ),Aφ=Fμ,AφfF_{f_{*}(\mu),A}^{\varphi}=F_{\mu,A}^{\varphi_{f}} where φf(x,z):=φ(f(x),z)\varphi_{f}(x,z):=\varphi(f(x),z).

(3) Let μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) be fim  over AA.

Claim.

For any θ(y1,,yn)(𝒰)\theta(y_{1},\ldots,y_{n})\in\mathcal{L}(\mathcal{U}),

(f(μ))(n)(θ(y1,,yn))=μ(n)(θ(f(x1),,f(xn))).(f_{*}(\mu))^{(n)}(\theta(y_{1},\ldots,y_{n}))=\mu^{(n)}(\theta(f(x_{1}),\ldots,f(x_{n}))).
Proof.

We prove the claim by induction on nn. The base case n=1n=1 is trivial. Assume the claim holds for nn. For 1k<ω1\leq k<\omega, let Gk:(𝒰x)k(𝒰y)kG_{k}:(\mathcal{U}^{x})^{k}\to(\mathcal{U}^{y})^{k} be the AA-definable map given by Gk(x1,,xk)=(f(x1),,f(xk))G_{k}(x_{1},\ldots,x_{k})=(f(x_{1}),\ldots,f(x_{k})). Note that GkG_{k} induces a pushforward from 𝔐x1,,xk(𝒰)\mathfrak{M}_{x_{1},\ldots,x_{k}}(\mathcal{U}) to 𝔐y1,,yk(𝒰)\mathfrak{M}_{y_{1},\ldots,y_{k}}(\mathcal{U}). Then our induction hypothesis says (f(μ))(n)=(Gn)(μ(n))(f_{*}(\mu))^{(n)}=\left(G_{n}\right)_{*}\left(\mu^{(n)}\right). Fix θ(y1,,yn+1)(𝒰)\theta(y_{1},\ldots,y_{n+1})\in\mathcal{L}(\mathcal{U}) and let

ψ(yn+1;y1,,yn):=θ(y1,,yn+1),\displaystyle\psi(y_{n+1};y_{1},\ldots,y_{n}):=\theta(y_{1},\ldots,y_{n+1}),
ψG(xn+1;x1,,xn):=θ(f(x1),,f(xn+1)).\displaystyle\psi_{G}(x_{n+1};x_{1},\ldots,x_{n}):=\theta(f(x_{1}),\ldots,f(x_{n+1})).

Let N𝒰N\prec\mathcal{U} be a small model containing AA and all relevant parameters. Then

(f(μ))(n+1)(θ(y1,,yn+1))=(f(μ)yn+1(f(μ))y1,,yn(n))(θ(y1,,yn+1))\displaystyle(f_{*}(\mu))^{(n+1)}(\theta(y_{1},\ldots,y_{n+1}))=\left(f_{*}(\mu)_{y_{n+1}}\otimes(f_{*}(\mu))^{(n)}_{y_{1},\ldots,y_{n}}\right)(\theta(y_{1},\ldots,y_{n+1}))
=Sy1,,yn(N)Ff(μ)ψd((f(μ))(n))=Sy1,,yn(N)Ff(μ)ψd((Gn)(μ(n)))\displaystyle=\int_{S_{y_{1},\ldots,y_{n}}(N)}F_{f_{*}(\mu)}^{\psi}\ \textrm{d}\left((f_{*}(\mu))^{(n)}\right)=\int_{S_{y_{1},\ldots,y_{n}}(N)}F_{f_{*}(\mu)}^{\psi}\ \textrm{d}\left(\left(G_{n}\right)_{*}\left(\mu^{(n)}\right)\right)
=Sx1,,xn(N)(Ff(μ)ψGn)dμ(n)=Sx1,,xn(N)FμψGdμ(n)\displaystyle=\int_{S_{x_{1},\ldots,x_{n}}(N)}\left(F_{f_{\ast}(\mu)}^{\psi}\circ G_{n}\right)\ \textrm{d}\mu^{(n)}=\int_{S_{x_{1},\ldots,x_{n}}(N)}F_{\mu}^{\psi_{G}}\ \textrm{d}\mu^{(n)}
=μ(n+1)(θ(f(x1),,f(xn+1))).\displaystyle=\mu^{(n+1)}(\theta(f(x_{1}),\ldots,f(x_{n+1}))).\qed

We now show that f(μ)f_{*}(\mu) is fim over AA. Fix a formula φ(y,z)\varphi(y,z) in yz\mathcal{L}_{yz}. Since μ\mu is fim, let (θn(x1,,xn))1n<ω(\theta_{n}(x_{1},\ldots,x_{n}))_{1\leq n<\omega} be a sequence of (A)\mathcal{L}(A)-formulas witnessing this for the formula φ(f(x),z)\varphi(f(x),z) as in Definition 3.5. To avoid “scope-of-quantifiers” confusion, we let w1,,wnw_{1},\ldots,w_{n} be new variables with wiw_{i} of the same sort as xix_{i} for each ii. For each 1n<ω1\leq n<\omega, we consider the (A)\mathcal{L}(A)-formula

γn(y1,,yn):=w1wn(θ(w1,,wn)1inf(wi)=yi).\gamma_{n}(y_{1},\ldots,y_{n}):=\exists w_{1}\ldots\exists w_{n}\left(\theta(w_{1},\ldots,w_{n})\wedge\bigwedge_{1\leq i\leq n}f(w_{i})=y_{i}\right).

Note that

(\ast) θ(x1,,xn)γn(f(x1),,f(xn)) for every nω.\theta(x_{1},\ldots,x_{n})\vdash\gamma_{n}(f(x_{1}),\ldots,f(x_{n}))\textrm{ for every }n\in\omega.

We will show that the formulas (γn(y1,,yn))1n<ω(\gamma_{n}(y_{1},\ldots,y_{n}))_{1\leq n<\omega} witness that f(μ)f_{*}(\mu) is fim over AA with respect to the formula φ(y,z)\varphi(y,z).

Fix ε>0\varepsilon>0. Then there exists some nεωn_{\varepsilon}\in\omega so that: for any nεk<ωn_{\varepsilon}\leq k<\omega and any d¯\bar{d} with θk(d¯)\models\theta_{k}(\bar{d}) we have

(\ast\ast) supb𝒰z|Av(d¯)(φ(f(x),b))μ(φ(f(x),b))|<ε.\sup_{b\in\mathcal{U}^{z}}|\operatorname{Av}(\bar{d})(\varphi(f(x),b))-\mu(\varphi(f(x),b))|<\varepsilon.

Now, suppose that c¯\bar{c} is such that 𝒰γk(c¯)\mathcal{U}\models\gamma_{k}(\bar{c}). Then, by definition of γk\gamma_{k}, there exists some e¯(𝒰x)k\bar{e}\in(\mathcal{U}^{x})^{k} such that θk(e¯)\models\theta_{k}(\bar{e}) and (f(e1),,f(ek))=(c1,,ck)(f(e_{1}),\ldots,f(e_{k}))=(c_{1},\ldots,c_{k}). Therefore,

supb𝒰z|Av(c¯)(φ(y,b))f(μ)(φ(y,b))|\displaystyle\sup_{b\in\mathcal{U}^{z}}|\operatorname{Av}(\overline{c})(\varphi(y,b))-f_{*}(\mu)(\varphi(y,b))|
=supb𝒰z|Av(c1,,cn)(φ(y,b))μ(φ(f(x),b))|\displaystyle=\sup_{b\in\mathcal{U}^{z}}|\operatorname{Av}(c_{1},\ldots,c_{n})(\varphi(y,b))-\mu(\varphi(f(x),b))|
=supb𝒰z|Av(f(e1),,f(en))(φ(y,b))μ(φ(f(x),b))|\displaystyle=\sup_{b\in\mathcal{U}^{z}}|\operatorname{Av}(f(e_{1}),\ldots,f(e_{n}))(\varphi(y,b))-\mu(\varphi(f(x),b))|
=supb𝒰z|Av(e1,,en)(φ(f(x),b))μ(φ(f(x),b))|<()ε.\displaystyle=\sup_{b\in\mathcal{U}^{z}}|\operatorname{Av}(e_{1},\ldots,e_{n})(\varphi(f(x),b))-\mu(\varphi(f(x),b))|\overset{(\ast\ast)}{<}\varepsilon.

Finally, by the Claim and ()(\ast) we have

limn((f(μ))(n)(γn(y1,,yn)))=limnμ(n)(γn(f(x1),,f(xn)))\displaystyle\lim_{n\to\infty}\left((f_{*}(\mu))^{(n)}(\gamma_{n}(y_{1},\ldots,y_{n}))\right)=\lim_{n\to\infty}\mu^{(n)}\left(\gamma_{n}(f(x_{1}),\ldots,f(x_{n}))\right)
limnμ(n)(θn(x1,,xn))=1.\displaystyle\geq\lim_{n\to\infty}\mu^{(n)}(\theta_{n}(x_{1},\ldots,x_{n}))=1.

We conclude f(μ)f_{*}(\mu) is fim.

(4) Immediate from the definitions. ∎

3.6. Fim and fsg groups

Throughout this section, G(x)G(x) will be a \emptyset-type-definable group. Let φ0(x)()\varphi_{0}(x)\in\mathcal{L}(\emptyset) be chosen for G(x)G(x) as in Section 2.2.

Definition 3.27.

[HPP08, Definition 4.1] A (\emptyset-)type-definable group GG is fsg (finitely satisfiable generics) if there is some pSG(𝒰)p\in S_{G}(\mathcal{U}) and small M𝒰M\prec\mathcal{U} such that for every gG(𝒰)g\in G(\mathcal{U}), gpg\cdot p is finitely satisfiable in MM.

We consider a natural generalization of generically stable groups (Section 2.3) to fim groups.

Definition 3.28.
  1. (1)

    We let 𝔐G(𝒰)\mathfrak{M}_{G}(\mathcal{U}) denote the (closed) set of all measures μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) supported on GG, i.e. with S(μ)SG(𝒰)S(\mu)\subseteq S_{G}(\mathcal{U}).

  2. (2)

    For μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) and gG(𝒰)g\in G(\mathcal{U}), we let gμ:=(g)(μ)g\cdot\mu:=\left(g\cdot-\right)_{\ast}(\mu) and μ1:=()1(μ)\mu^{{}^{-}1}:=\left({}^{-1}\right)_{\ast}(\mu) (where, as in Section 2.2, we view ,1\cdot,^{-1} as globally defined functions whose restrictions to GG give the group operations). As gg\cdot- and -1 are definable bijections on φ0(𝒰)\varphi_{0}(\mathcal{U}), we get S(gμ)=gS(μ)S(g\cdot\mu)=g\cdot S(\mu) and S(μ1)={p1:pS(μ)}S(\mu^{-1})=\{p^{-1}:p\in S(\mu)\} (by Proposition 3.26(4)). In particular, gμ,μ1𝔐G(𝒰)g\cdot\mu,\mu^{-1}\in\mathfrak{M}_{G}(\mathcal{U}). We define the right action of GG on 𝔐G(𝒰)\mathfrak{M}_{G}(\mathcal{U}) analogously.

Definition 3.29.

We will say that a (\emptyset-)type-definable group G(x)G(x) is fim if there exists a right GG-invariant fim  measure μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}), i.e. μg=μ\mu\cdot g=\mu for all gG(𝒰)g\in G(\mathcal{U}).

Remark 3.30.
  1. (1)

    In any theory, if GG is fim then it is both definably amenable and fsg.

  2. (2)

    If TT is NIP and GG is fsg, then it is fim.

Proof.

(1) Indeed, let μ\mu witness that GG is fim. In particular, μ\mu is finitely satisfiable in some small M𝒰M\prec\mathcal{U} (Remark 3.6), and gμ=μg\cdot\mu=\mu for all gG(𝒰)g\in G(\mathcal{U}). Take any pS(μ)p\in S(\mu). Then pSG(𝒰)p\in S_{G}(\mathcal{U}), and also gpS(μ)g\cdot p\in S(\mu), hence finitely satisfiable in MM, for all gG(𝒰)g\in G(\mathcal{U}).

(2) By [HPS13, Remark 4.4], and Remark 3.6. ∎

Problem 3.31.

Does fsg imply definable amenability without assuming NIP? Do there exist fsg (and definably amenable) groups that are not fim?

The following is a simultaneous generalization of Fact 2.5 from types to measures in arbitrary theories, and of the previously known case for measures under the NIP assumption [HPS13, Theorem 4.3].

Proposition 3.32.

Suppose that G(x)G(x) is a \emptyset-type-definable fim group, witnessed by a right-GG-invariant fim measure μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}). Then we have:

  1. (1)

    μ=μ1\mu=\mu^{-1};

  2. (2)

    μ\mu is left GG-invariant;

  3. (3)

    μ\mu is the unique left GG-invariant measure in 𝔐G(𝒰)\mathfrak{M}_{G}(\mathcal{U});

  4. (4)

    μ\mu is the unique right GG-invariant measure in 𝔐G(𝒰)\mathfrak{M}_{G}(\mathcal{U}).

Proof.

(1) Fix a formula φ(x)x(𝒰)\varphi(x)\in\mathcal{L}_{x}(\mathcal{U}). Let M𝒰M\prec\mathcal{U} be a small model such that μ\mu is MM-invariant and MM contains the parameters of φ\varphi. As μ\mu is fim  over MM, μ1\mu^{-1} is also fim over MM by Proposition 3.26(3).

Suppose q(x)S(μx|M)q(x)\in S(\mu_{x}|_{M}) (so q(x)G(x)q(x)\vdash G(x)). Then, for (φ)(y,x)=φ(xy)(\varphi^{\prime})^{*}(y,x)=\varphi(x\cdot y) (in the notation of Section 3.2) we have (for any bq(x)b\models q(x)):

Fμy1(φ)(q)=μy1(φ(by))=μy(φ(by1))=μy(φ((yb1)1))\displaystyle F_{\mu_{y}^{-1}}^{(\varphi^{\prime})^{*}}(q)=\mu_{y}^{-1}(\varphi(b\cdot y))=\mu_{y}(\varphi(b\cdot y^{-1}))=\mu_{y}(\varphi((y\cdot b^{-1})^{-1}))
=()μy(φ(y1))=μy1(φ(y)),\displaystyle\overset{(*)}{=}\mu_{y}(\varphi(y^{-1}))=\mu_{y}^{-1}(\varphi(y)),

where ()(*) follows by right GG-invariance of μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) applied to the formula ψ(y):=φ(y1)\psi(y):=\varphi(y^{-1}).

Using this and that fim measures commute with all Borel definable measures (Fact 3.7) we have:

μμ1(φ(x))=μxμy1(φ(xy))=μy1μx(φ(xy))\displaystyle\mu*\mu^{-1}(\varphi(x))=\mu_{x}\otimes\mu^{-1}_{y}(\varphi(x\cdot y))=\mu^{-1}_{y}\otimes\mu_{x}(\varphi(x\cdot y))
=Sx(M)Fμy1(φ)dμx=S(μx|M)Fμy1(φ)dμx=Sx(M)μy1(φ(y))dμx\displaystyle=\int_{S_{x}(M)}F_{\mu^{-1}_{y}}^{(\varphi^{\prime})^{*}}\textrm{d}\mu_{x}=\int_{S(\mu_{x}|_{M})}F_{\mu^{-1}_{y}}^{(\varphi^{\prime})^{*}}\textrm{d}\mu_{x}=\int_{S_{x}(M)}\mu_{y}^{-1}(\varphi(y))\textrm{d}\mu_{x}
=μy1(φ(y))=μx1(φ(x)).\displaystyle=\mu_{y}^{-1}(\varphi(y))=\mu_{x}^{-1}(\varphi(x)).

Similarly, for φ(x,y)=φ(xy)\varphi^{\prime}(x,y)=\varphi(x\cdot y) and any q(y)S(μy1|M)q(y)\in S(\mu^{-1}_{y}|_{M}) (so q(y)G(y)q(y)\vdash G(y)), by right GG-invariance of μ\mu we have Fμxφ(q)=μ(φ(x))F_{\mu_{x}}^{\varphi^{\prime}}(q)=\mu(\varphi(x)). Then

μμ1(φ(x))=μxμy1(φ(xy))=Sy(M)Fμxφd(μy1)\displaystyle\mu*\mu^{-1}(\varphi(x))=\mu_{x}\otimes\mu^{-1}_{y}(\varphi(x\cdot y))=\int_{S_{y}(M)}F_{\mu_{x}}^{\varphi^{\prime}}\textrm{d}\left(\mu^{-1}_{y}\right)
=Sy(M)μ(φ(x))d(μy1)=μ(φ(x)).\displaystyle=\int_{S_{y}(M)}\mu(\varphi(x))\textrm{d}\left(\mu^{-1}_{y}\right)=\mu(\varphi(x)).

We conclude that μ(φ(x))=μ1(φ(x))\mu(\varphi(x))=\mu^{-1}(\varphi(x)).

(2) For any aG(𝒰)a\in G(\mathcal{U}) and φ(x)(𝒰)\varphi(x)\in\mathcal{L}(\mathcal{U}), using (1) and right GG-invariance of μ\mu for the formula ψ(x):=φ(x1)\psi(x):=\varphi(x^{-1}) we have:

μ(φ(ax))=μ1(φ(ax))=μ(φ(ax1))=μ(φ((xa1)1))\displaystyle\mu(\varphi(a\cdot x))=\mu^{-1}(\varphi(a\cdot x))=\mu(\varphi(a\cdot x^{-1}))=\mu(\varphi((x\cdot a^{-1})^{-1}))
=μ(φ(x1))=μ1(φ(x))=μ(φ(x)).\displaystyle=\mu(\varphi(x^{-1}))=\mu^{-1}(\varphi(x))=\mu(\varphi(x)).

(3) Suppose that ν𝔐G(𝒰)\nu\in\mathfrak{M}_{G}(\mathcal{U}) is left GG-invariant, and let φ(x)(𝒰)\varphi(x)\in\mathcal{L}(\mathcal{U}) be arbitrary. Let M𝒰M\prec\mathcal{U} be a small model such that μ\mu is invariant over MM and MM contains the parameters of φ\varphi.

As in (1), for any q(y)S(νy|M)q(y)\in S(\nu_{y}|_{M}) (so q(y)G(y)q(y)\vdash G(y)), by right GG-invariance of μ\mu we have Fμxφ(q)=μ(φ(x))F_{\mu_{x}}^{\varphi^{\prime}}(q)=\mu(\varphi(x)). Hence

(*) Sy(M)Fμxφdνy=S(ν|M)Fμxφdνy=S(ν|M)μ(φ(x))dνy=Sy(M)μ(φ(x))dνy.\int_{S_{y}(M)}F_{\mu_{x}}^{\varphi^{\prime}}\textrm{d}\nu_{y}=\int_{S(\nu|_{M})}F_{\mu_{x}}^{\varphi^{\prime}}\textrm{d}\nu_{y}=\int_{S(\nu|_{M})}\mu(\varphi(x))\textrm{d}\nu_{y}=\int_{S_{y}(M)}\mu(\varphi(x))\textrm{d}\nu_{y}.

Second, for any qS(μx|M)q\in S(\mu_{x}|_{M}) (so q(x)G(x)q(x)\vdash G(x)), by left GG-invariance of ν\nu we have (for any bqb\models q):

(†) Fνy(φ)(q)=ν(φ(by))=ν(φ(y)).F_{\nu_{y}}^{(\varphi^{\prime})^{*}}(q)=\nu(\varphi(b\cdot y))=\nu(\varphi(y)).

So the map Fνy(φ):S(μ|M)[0,1]F_{\nu_{y}}^{(\varphi^{\prime})^{*}}:S(\mu|_{M})\to[0,1] is constant on the support of μ\mu, hence Borel. As μ\mu is fim, the proof of [CGH23a, Proposition 5.15] implies

(‡) S(μ|M)Fνy(φ)dμx=Sy(M)Fμxφdνy,\int_{S(\mu|_{M})}F_{\nu_{y}}^{(\varphi^{\prime})^{*}}\textrm{d}\mu_{x}=\int_{S_{y}(M)}F_{\mu_{x}}^{\varphi^{\prime}}\textrm{d}\nu_{y},

where on the left we view μ\mu as a regular Borel probability measure restricted to the compact set S(μ|M)S(\mu|_{M}). Then

μ(φ(x))=Sy(M)μ(φ(x))dνy=()Sy(M)Fμxφ𝑑νy\displaystyle\mu(\varphi(x))=\int_{S_{y}(M)}\mu(\varphi(x))\textrm{d}\nu_{y}\overset{(*)}{=}\int_{S_{y}(M)}F_{\mu_{x}}^{\varphi^{\prime}}d\nu_{y}
=()S(μ|M)Fνy(φ)dμx=()S(μ|M)ν(φ(y))dμx=Sx(M)ν(φ(y))dμx=ν(φ(y)).\displaystyle\overset{({\ddagger})}{=}\int_{S(\mu|_{M})}F_{\nu_{y}}^{(\varphi^{\prime})^{*}}\textrm{d}\mu_{x}\overset{({\dagger})}{=}\int_{S(\mu|_{M})}\nu(\varphi(y))\textrm{d}\mu_{x}=\int_{S_{x}(M)}\nu(\varphi(y))\textrm{d}\mu_{x}=\nu(\varphi(y)).

(4) Let ν𝔐G(𝒰)\nu\in\mathfrak{M}_{G}(\mathcal{U}) be right GG-invariant, φ(x)(𝒰)\varphi(x)\in\mathcal{L}(\mathcal{U}) and M𝒰M\prec\mathcal{U} containing the parameters of φ\varphi and such that μ\mu is MM-invariant.

As in (3), the map Fνxφ:Sy(M)[0,1]F_{\nu_{x}}^{\varphi^{\prime}}:S_{y}(M)\to[0,1] is (μy|M)(\mu_{y}|_{M})-measurable since it is constant on the support by right GG-invariance of ν\nu. As μ\mu is fim we can apply [CGH23a, Proposition 5.15] again to get

Sy(M)Fνxφdμy=Sx(M)Fμy(φ)dνx.\int_{S_{y}(M)}F_{\nu_{x}}^{\varphi^{\prime}}\textrm{d}\mu_{y}=\int_{S_{x}(M)}F_{\mu_{y}}^{(\varphi^{\prime})^{*}}\textrm{d}\nu_{x}.

As μ\mu is left GG-invariant by (2), for any qS(νx|M)q\in S(\nu_{x}|_{M}), we have (for bqb\models q):

Fμy(φ)(q)=μy(φ(by))=μ(φ(y)).F_{\mu_{y}}^{(\varphi^{\prime})^{*}}(q)=\mu_{y}(\varphi(b\cdot y))=\mu(\varphi(y)).

Combining we get

ν(φ(x))=Sy(M)ν(φ(x))dμy=Sy(M)Fνxφdμy\displaystyle\nu(\varphi(x))=\int_{S_{y}(M)}\nu(\varphi(x))\textrm{d}\mu_{y}=\int_{S_{y}(M)}F_{\nu_{x}}^{\varphi^{\prime}}\textrm{d}\mu_{y}
=Sx(M)Fμy(φ)dνx=Sx(M)μ(φ(y))dνx=μ(φ(y)).\displaystyle=\int_{S_{x}(M)}F_{\mu_{y}}^{(\varphi^{\prime})^{*}}\textrm{d}\nu_{x}=\int_{S_{x}(M)}\mu(\varphi(y))\textrm{d}\nu_{x}=\mu(\varphi(y)).\qed

Proposition 3.32 and its symmetric version with “left” swapped with “right” (obtained by a “symmetric” proof) yield:

Corollary 3.33.

An (\emptyset)-type-definable group G(x)G(x) is fim if and only if there exists a left GG-invariant fim measure μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}).

3.7. Idempotent fim measures and generic transitivity

Let G(x)G(x) be a \emptyset-type-definable group, and we are in the same setting and notation as in Section 2.2. For a measure μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}), we let

Stab(μ):={gG(𝒰):μg=g}\operatorname{Stab}(\mu):=\{g\in G(\mathcal{U}):\mu\cdot g=g\}

denote the right-stabilizer of μ\mu.

Fact 3.34.

When μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is a measure definable over M𝒰M\prec\mathcal{U}, then Stab(μ)\operatorname{Stab}(\mu) is an MM-type-definable subgroup of G(𝒰)G(\mathcal{U}) (see e.g. [CG22, Proposition 5.3]).

Definition 3.35.

Suppose that μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is Borel-definable. Then for any measure ν𝔐G(𝒰)\nu\in\mathfrak{M}_{G}(\mathcal{U}), the (definable) convolution of μ\mu and ν\nu, denoted μν\mu*\nu, is the unique measure in 𝔐G(𝒰)\mathfrak{M}_{G}(\mathcal{U}) such that for any formula φ(x)(𝒰)\varphi(x)\in\mathcal{L}(\mathcal{U}),

(μν)(φ(x))=(μν)(φ(xy)).(\mu*\nu)(\varphi(x))=(\mu\otimes\nu)(\varphi(x\cdot y)).

We say that μ\mu is idempotent if μμ=μ\mu*\mu=\mu.

When TT is NIP, it suffices to assume that μ\mu is invariant (under Aut(𝕄/M)\operatorname{Aut}(\mathbb{M}/M) for some small model MM), as then μ\mu is automatically Borel-definable ([Sim15]). We refer to [CG22, Section 3] for a detailed discussion of when convolution is well-defined.

We now consider the main question investigated in [CG22, CG23] in the case of measures (generalizing from types in Section 2.4).

Again, we let H:=Stab(μ)H:=\operatorname{Stab}(\mu).

Definition 3.36.

For the rest of the section, we let f:(𝒰x)2(𝒰x)2f:\left(\mathcal{U}^{x}\right)^{2}\to\left(\mathcal{U}^{x}\right)^{2} be the (\emptyset-definable) map f(x1,x0)=(x1x0,x0)f(x_{1},x_{0})=(x_{1}\cdot x_{0},x_{0}) (where \cdot is viewed as a globally defined function whose restriction to GG defines the group operation, see Section 2.2).

Proposition 3.37.

Let μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) be an idempotent fim measure. Then the following are equivalent:

  1. (1)

    μ𝔐H(𝒰)\mu\in\mathfrak{M}_{H}(\mathcal{U});

  2. (2)

    μ(2)=f(μ(2))\mu^{(2)}=f_{\ast}\left(\mu^{(2)}\right);

  3. (3)

    μp=f(μp)\mu\otimes p=f_{\ast}(\mu\otimes p) for every pS(μ)p\in S(\mu).

Proof.

(2) \Leftrightarrow (3) This equivalence only uses that μ\mu is a definable measure.

By the definition of ff we have f(μ(2))(φ(x1,x0))=μ(2)(φ~(x1,x0))f_{\ast}\left(\mu^{(2)}\right)\left(\varphi(x_{1},x_{0})\right)=\mu^{(2)}\left(\tilde{\varphi}(x_{1},x_{0})\right) for all φ(x1,x0)x1,x0(𝒰)\varphi(x_{1},x_{0})\in\mathcal{L}_{x_{1},x_{0}}\left(\mathcal{U}\right), where φ~(x1,x0):=φ(x1x0,x0)\tilde{\varphi}(x_{1},x_{0}):=\varphi(x_{1}\cdot x_{0},x_{0}). By definition of μ(2)\mu^{(2)} we have: μ(2)f(μ(2))\mu^{(2)}\neq f_{\ast}\left(\mu^{(2)}\right) if and only if there exists a formula φ(x1,x0)x1,x0(𝒰)\varphi(x_{1},x_{0})\in\mathcal{L}_{x_{1},x_{0}}(\mathcal{U}) and a small model M𝒰M\prec\mathcal{U} so that MM contains the parameters of φ\varphi and μ\mu is MM-definable, such that

S(μ|M)|Fμ,Mφ~(p)Fμ,Mφ(p)|dμM>0,\int_{S(\mu|_{M})}\left\lvert F_{\mu,M}^{\tilde{\varphi}}(p)-F^{\varphi}_{\mu,M}(p)\right\rvert\textrm{d}\mu_{M}>0,

where μ|M𝔐x0(M)\mu|_{M}\in\mathfrak{M}_{x_{0}}(M) is the restriction of μ\mu to MM, μM\mu_{M} is the restriction μ|M\mu|_{M} viewed as a regular Borel measure on Sx0(M)S_{x_{0}}(M). By definability of μ\mu, the function pS(μ|M)|Fμ,Mφ~(p)Fμ,Mφ(p)|[0,1]p\in S(\mu|_{M})\mapsto\left\lvert F_{\mu,M}^{\tilde{\varphi}}(p)-F^{\varphi}_{\mu,M}(p)\right\rvert\in[0,1] is continuous (and non-negative), hence the integral is >0>0 if and only if |Fμ,Mφ~(p)Fμ,Mφ(p)|>0\left\lvert F_{\mu,M}^{\tilde{\varphi}}(p)-F^{\varphi}_{\mu,M}(p)\right\rvert>0 for some pS(μ|M)p\in S(\mu|_{M}), that is μ(φ(x1,b))μ(φ(x1b,b))\mu\left(\varphi(x_{1},b)\right)\neq\mu(\varphi(x_{1}\cdot b,b)) for some pS(μ)p\in S(\mu) and bp|Mb\models p|_{M}, that is (μp)(φ(x1,x0))f(μp)(φ(x1,x0))\left(\mu\otimes p\right)\left(\varphi(x_{1},x_{0})\right)\neq f_{\ast}\left(\mu\otimes p\right)\left(\varphi(x_{1},x_{0})\right) for some pS(μ)p\in S(\mu).

(3) \Rightarrow (1) Fix pS(μ)p\in S(\mu), let M𝒰M\prec\mathcal{U} be such that μ\mu is MM-invariant, and let ap|Ma\models p|_{M}.

Claim 3.

We have μ|Ma=(μa)|Ma\mu|_{Ma}=\left(\mu\cdot a\right)|_{Ma}.

Proof.

Any ψ(x)x(Ma)\psi(x)\in\mathcal{L}_{x}(Ma) is of the form φ(x,a)\varphi(x,a) for some φ(x,y)xy(M)\varphi(x,y)\in\mathcal{L}_{xy}(M). By (3) we have:

μ(φ(x,a))=(μp)(φ(x,y))=f(μp)(φ(x,y))\displaystyle\mu(\varphi(x,a))=(\mu\otimes p)(\varphi(x,y))=f_{\ast}(\mu\otimes p)(\varphi(x,y))
=(μp)(φ(xy,y))=μ(φ(xa,a)).\displaystyle=(\mu\otimes p)(\varphi(x\cdot y,y))=\mu(\varphi(x\cdot a,a)).

Therefore,

(μa)(ψ(x))=μ(ψ(xa))=μ(φ(xa,a))=μ(φ(x,a))=μ(ψ(x)).\displaystyle\left(\mu\cdot a\right)(\psi(x))=\mu(\psi(x\cdot a))=\mu(\varphi(x\cdot a,a))=\mu(\varphi(x,a))=\mu(\psi(x)).

Claim 4.

The measure μa\mu\cdot a is (Ma)(Ma)-definable.

Proof.

Since μ\mu is MM-definable, it is also MaMa-definable. Consider the MaMa-definable map g:xxag:x\mapsto x\cdot a. Note that g(μ)=μag_{\ast}(\mu)=\mu\cdot a and so by Proposition 3.26, μa\mu\cdot a is also MaMa-definable. ∎

Hence, by Proposition 3.11 over MaMa, we get μ=μa\mu=\mu\cdot a. That is, aStab(μ)a\in\operatorname{Stab}(\mu), so p|M(x)H(x)p|_{M}(x)\vdash H(x).

(1) \Rightarrow (3) Let φ(x,y)(𝒰)\varphi(x,y)\in\mathcal{L}(\mathcal{U}) be arbitrary, and M𝒰M\prec\mathcal{U} contain its parameters such that μ\mu is MM-invariant. The measure μ\mu is right HH-invariant, and given any pS(μ)p\in S(\mu) we have pSH(𝒰)p\in S_{H}(\mathcal{U}) by (1). So for ap|Ma\models p|_{M} we have aHa\in H, hence μp(φ(x,y))=μ(φ(x,a))=μ(φ(xa,a))=f(μp)(φ(x,y))\mu\otimes p(\varphi(x,y))=\mu(\varphi(x,a))=\mu(\varphi(x\cdot a,a))=f_{\ast}(\mu\otimes p)(\varphi(x,y)). ∎

Definition 3.38.

We say that an idempotent fim measure μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is generically transitive if it satisfies any of the equivalent conditions in Proposition 3.37.

In particular, a type is generically transitive in the sense of Definition 2.13 if and only if, viewed as a Keisler measure, it is generically transitive.

The following is a generalization of Remark 2.11:

Remark 3.39.

Assume μ\mu is fim  and μ𝔐H(𝒰)\mu\in\mathfrak{M}_{H}(\mathcal{U}). Then:

  1. (1)

    HH is a fim group (Definition 3.29), hence μ\mu is both the unique left-invariant and the unique right-invariant measure supported on HH (by Proposition 3.32);

  2. (2)

    HH is the smallest among all type-definable subgroups HH^{\prime} of GG with μ𝔐H(𝒰)\mu\in\mathfrak{M}_{H^{\prime}}(\mathcal{U});

  3. (3)

    HH is both the left and the right stabilizer of μ\mu in GG.

Proof.

(1) HH is a fim group, witnessed by the right HH-invariant fim measure μ𝔐H(𝒰)\mu\in\mathfrak{M}_{H}(\mathcal{U}).

(2) HH is type-definable by Fact 3.34.

For any type-definable HG(𝒰)H^{\prime}\leq G(\mathcal{U}) with S(μ)SH(𝒰)S(\mu)\subseteq S_{H^{\prime}}(\mathcal{U}), the group H′′:=HHH^{\prime\prime}:=H\cap H^{\prime} is type-definable with S(μ)SH′′(𝒰)S(\mu)\subseteq S_{H^{\prime\prime}}(\mathcal{U}) and H′′HH^{\prime\prime}\leq H. If the index is 2\geq 2, we have some gHg\in H with H′′(H′′g)=H^{\prime\prime}\cap\left(H^{\prime\prime}\cdot g\right)=\emptyset, and S(μ)SH′′(𝒰),S(μg)SH′′g(𝒰)S(\mu)\subseteq S_{H^{\prime\prime}}(\mathcal{U}),S\left(\mu\cdot g\right)\subseteq S_{H^{\prime\prime}\cdot g}(\mathcal{U}), so S(μ)S(μg)=S(\mu)\cap S(\mu\cdot g)=\emptyset, so μμg\mu\neq\mu\cdot g — a contradiction. So H′′=HH^{\prime\prime}=H, and HHH\subseteq H^{\prime}.

(3) Let HH_{\ell} be the left stabilizer of μ\mu. Then HH_{\ell} is type-definable by a symmetric version of Fact 3.34. By (1), μ\mu is left HH-invariant, so we have HHH\subseteq H_{\ell}, and so S(μ)SH(𝒰)S(\mu)\subseteq S_{H_{\ell}}(\mathcal{U}) is left HH_{\ell}-invariant. By a symmetric argument as in (2), we conclude H=HH=H_{\ell}. ∎

Example 3.40.

If GG^{\prime} is a fim type-definable subgroup of GG, witnessed by a GG^{\prime}-invariant fim measure μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G^{\prime}}(\mathcal{U}), then μ\mu is obviously idempotent and generically transitive.

Analogously to the case of types (Section 2.4), the following is our main question in the case of measures:

Problem 3.41.

Assume that μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is fim and idempotent. Is it true that then μ\mu is generically transitive? Assuming TT is NIP?

A positive answer for stable TT is given in [CG22], see Section 3.10 for a discussion.

By symmetric versions of the above considerations, one easily gets:

Remark 3.42.

Section 3.7 remains valid if one swaps “left” with “right” (including left and right stabilizers) and replaces ff by f(x1,x0)=(x0x1,x0)f(x_{1},x_{0})=(x_{0}\cdot x_{1},x_{0}).

3.8. Idempotent fim measures in abelian groups

Finally, we have all of the ingredients to adapt the proof in Section 2.5 from types to measures and give a positive answer to Problem 3.41 for abelian groups in arbitrary theories. Let G(x)G(x) and ff be as in the previous section.

First note the idempotency of a fim measure μ\mu can be iterated along a Morley sequence in μ\mu:

Lemma 3.43.

Let μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) and suppose that μ\mu is fim and idempotent. Then for any formula ψ(x)(𝒰)\psi(x)\in\mathcal{L}(\mathcal{U}) and kωk\in\omega we have

μ(ψ(x))=μ(k)(ψ(x1xk)).\mu(\psi(x))=\mu^{(k)}(\psi(x_{1}\cdot\ldots\cdot x_{k})).
Proof.

By induction on kk, the base case k=2k=2 is the assumption on μ\mu. So assume that for any γ(x)(𝒰)\gamma(x)\in\mathcal{L}(\mathcal{U}) and k1\ell\leq k-1 we have

(a) μx1,,x()(γ(x1x))=μ(γ(x)).\displaystyle\mu^{(\ell)}_{x_{1},\ldots,x_{\ell}}(\gamma(x_{1}\cdot\ldots\cdot x_{\ell}))=\mu(\gamma(x)).

Fix ψ(x)(𝒰)\psi(x)\in\mathcal{L}(\mathcal{U}), and choose a small M𝒰M\prec\mathcal{U} such that μ\mu is MM-invariant and MM contains the parameters of ψ\psi. Let θ(x1,,xk1;xk):=ψ(x1xk)\theta(x_{1},\ldots,x_{k-1};x_{k}):=\psi(x_{1}\cdot\ldots\cdot x_{k}) and φ(x1;xk):=ψ(x1xk)\varphi(x_{1};x_{k}):=\psi(x_{1}\cdot x_{k}). For any qSxk(M)q\in S_{x_{k}}(M) (and any bqb\models q) we have

(b) Fμx1,,xk1(k1)θ(q)=μ(k1)(ψ(x1xk1b))=(a)μ(ψ(x1b))=Fμx1φ(q).\displaystyle F^{\theta}_{\mu^{(k-1)}_{x_{1},\ldots,x_{k-1}}}(q)=\mu^{(k-1)}\left(\psi(x_{1}\cdot\ldots\cdot x_{k-1}\cdot b)\right)\overset{\textrm{(a)}}{=}\mu(\psi(x_{1}\cdot b))=F^{\varphi}_{\mu_{x_{1}}}(q).

Then, using that μ\mu is fim and fim measures commute with Borel definable measures,

μ(k)(ψ(x1xk))=μxkμx1,,xk1(k1)(ψ(x1xk))\displaystyle\mu^{(k)}\left(\psi(x_{1}\cdot\ldots\cdot x_{k})\right)=\mu_{x_{k}}\otimes\mu^{(k-1)}_{x_{1},\ldots,x_{k-1}}(\psi(x_{1}\cdot\ldots\cdot x_{k}))
=μx1,,xk1(k1)μxk(ψ(x1xk))=Sxk(M)Fμx1,,xk1(k1)θdμxk\displaystyle=\mu^{(k-1)}_{x_{1},\ldots,x_{k-1}}\otimes\mu_{x_{k}}(\psi(x_{1}\cdot\ldots\cdot x_{k}))=\int_{S_{x_{k}}(M)}F^{\theta}_{\mu^{(k-1)}_{x_{1},\ldots,x_{k-1}}}\textrm{d}\mu_{x_{k}}
=(b)Sxk(M)Fμx1φdμxk=μx1μxk(φ(x1,xk))=μx1μxk(ψ(x1xk))=(a)μ(ψ(x)).\displaystyle\overset{\textrm{(b)}}{=}\int_{S_{x_{k}}(M)}F^{\varphi}_{\mu_{x_{1}}}\textrm{d}\mu_{x_{k}}=\mu_{x_{1}}\otimes\mu_{x_{k}}\left(\varphi(x_{1},x_{k})\right)=\mu_{x_{1}}\otimes\mu_{x_{k}}\left(\psi(x_{1}\cdot x_{k})\right)\overset{\textrm{(a)}}{=}\mu(\psi(x)).\qed
Lemma 3.44.

Assume that μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is fim and idempotent. Let k2k\geq 2 be arbitrary and let g:(𝒰x)k(𝒰x)k+1g:(\mathcal{U}^{x})^{k}\to(\mathcal{U}^{x})^{k+1} be the definable map

g(x1,,xk)=(x1xk,x1,,xk),g(x_{1},\ldots,x_{k})=(x_{1}\cdot\ldots\cdot x_{k},x_{1},\ldots,x_{k}),

where “\cdot” is viewed as a globally defined function whose restriction to GG defines the group operation (see Section 2.2). Let λk(y,x1,,xk):=g(μ(k))\lambda_{k}(y,x_{1},\ldots,x_{k}):=g_{\ast}\left(\mu^{(k)}\right). Then:

  1. (1)

    λk|x1,,xk=μ(k)\lambda_{k}|_{x_{1},\ldots,x_{k}}=\mu^{(k)};

  2. (2)

    if GG is abelian, then λk|y,xi=f(μ(2)(y,xi))\lambda_{k}|_{y,x_{i}}=f_{\ast}\left(\mu^{(2)}(y,x_{i})\right) for every 1ik1\leq i\leq k.

Proof.

(1) By definition of gg.

(2) Let φ(y,x)(𝒰)\varphi(y,x)\in\mathcal{L}(\mathcal{U}) be arbitrary, and let M𝒰M\prec\mathcal{U} be a small model containing its parameters, and so that μ\mu is MM-invariant. We let

θ(x1,,xi1,xi+1,,xk;xi):=φ(x1xk,xi),ψ(x;y):=φ(xy,y),\displaystyle\theta(x_{1},\ldots,x_{i-1},x_{i+1},\ldots,x_{k};x_{i}):=\varphi(x_{1}\cdot\ldots\cdot x_{k},x_{i}),\ \psi(x;y):=\varphi(x\cdot y,y),
μx^i(k1):=μxk1μxi+1μxi1μx1\displaystyle\mu^{(k-1)}_{\hat{x}_{i}}:=\mu_{x_{k-1}}\otimes\ldots\otimes\mu_{x_{i+1}}\otimes\mu_{x_{i-1}}\otimes\ldots\otimes\mu_{x_{1}}

Using that GG is abelian and Lemma 3.43, for any qSxi(M)q\in S_{x_{i}}(M) and bqb\models q we have (renaming the variables when necessary):

Fμx^i(k1)θ(q)=μx^i(k1)(φ(x1xi1bxi+1xk,b))\displaystyle F_{\mu^{(k-1)}_{\hat{x}_{i}}}^{\theta}(q)=\mu_{\hat{x}_{i}}^{(k-1)}(\varphi(x_{1}\cdot\ldots\cdot x_{i-1}\cdot b\cdot x_{i+1}\cdot\ldots\cdot x_{k},b))
=μx^i(k1)(φ(x1xi1xi+1xkb,b))\displaystyle=\mu_{\hat{x}_{i}}^{(k-1)}(\varphi(x_{1}\cdot\ldots\cdot x_{i-1}\cdot x_{i+1}\cdot\ldots\cdot x_{k}\cdot b,b))
=μx(φ(xb,b))=Fμxψ(q).\displaystyle=\mu_{x}(\varphi(x\cdot b,b))=F_{\mu_{x}}^{\psi}(q).

Hence, since GG is abelian, μ\mu is fim, and fim  measures commute with Borel definable measures, we get

λk(φ(y,xi))=μx1,,xk(k)(φ(x1xk,xi))\displaystyle\lambda_{k}(\varphi(y,x_{i}))=\mu_{x_{1},\ldots,x_{k}}^{(k)}(\varphi(x_{1}\cdot\ldots\cdot x_{k},x_{i}))
=μx^i(k1)μxi(φ(x1xi1xi+1xkxi,xi))\displaystyle=\mu_{\hat{x}_{i}}^{(k-1)}\otimes\mu_{x_{i}}(\varphi(x_{1}\cdot\ldots\cdot x_{i-1}\cdot x_{i+1}\ldots\cdot x_{k}\cdot x_{i},x_{i}))
=Sxi(M)Fμx^i(k1)θdμxi=Sxi(M)Fμxψdμxi\displaystyle=\int_{S_{x_{i}}(M)}F^{\theta}_{\mu_{\hat{x}_{i}}^{(k-1)}}\textrm{d}\mu_{x_{i}}=\int_{S_{x_{i}}(M)}F^{\psi}_{\mu_{x}}\textrm{d}\mu_{x_{i}}
=μxμxi(φ(xxi,xi))=μxμy(φ(xy,y))=f(μ(2))(φ(x,y)).\displaystyle=\mu_{x}\otimes\mu_{x_{i}}(\varphi(x\cdot x_{i},x_{i}))=\mu_{x}\otimes\mu_{y}(\varphi(x\cdot y,y))=f_{*}\left(\mu^{(2)}\right)(\varphi(x,y))\qed.

Finally, using results about the randomization (Theorem 3.13) and Lemma 3.44, we can show generic transitivity in the abelian case:

Theorem 3.45.

Assume that G(x)G(x) is an abelian type-definable group and μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is fim and idempotent. Then μ\mu is generically transitive.

Proof.

By Proposition 3.37, it suffices to show that μ(2)=f(μ(2))\mu^{(2)}=f_{\ast}\left(\mu^{(2)}\right). Assume not, say

|μ(2)(φ(x1,x2))f(μ(2))(φ(x1,x2))|=ε0\left\lvert\mu^{(2)}(\varphi(x_{1},x_{2}))-f_{\ast}\left(\mu^{(2)}\right)(\varphi(x_{1},x_{2}))\right\rvert=\varepsilon_{0}

for some φ(x1,x2)(𝒰)\varphi(x_{1},x_{2})\in\mathcal{L}(\mathcal{U}) and some ε0>0\varepsilon_{0}>0. Let M𝒰M\prec\mathcal{U} be a small model containing the parameters of φ\varphi, and so that μ\mu is invariant over MM. Let nn be as given by the moreover part of Theorem 3.13 for μ,φ,ε0\mu,\varphi,\varepsilon_{0}. Fix any k>nk>n, and consider the definable map g:(𝒰x)k(𝒰x)k+1g:(\mathcal{U}^{x})^{k}\to(\mathcal{U}^{x})^{k+1} given by g(x1,,xk)=(x1xk,x1,,xk)g(x_{1},\ldots,x_{k})=(x_{1}\cdot\ldots\cdot x_{k},x_{1},\ldots,x_{k}). Then gg induces a continuous map from S𝐱(𝒰)S_{\mathbf{x}}(\mathcal{U}) to S𝐱y(𝒰)S_{\mathbf{x}y}(\mathcal{U}), where 𝐱=(xi:iω)\mathbf{x}=(x_{i}:i\in\omega) and we let λ𝔐𝐱y(𝒰)\lambda\in\mathfrak{M}_{\mathbf{x}y}(\mathcal{U}) be defined by λ:=g(μ(ω))\lambda:=g_{\ast}(\mu^{(\omega)}). That is, for every mωm\in\omega and every φ(x1,,xm,y)(𝒰)\varphi(x_{1},\ldots,x_{m},y)\in\mathcal{L}(\mathcal{U}) we have

λ(φ(x1,,xm,y))=μ(ω)(φ(x1,,xm,x1xk)).\lambda(\varphi(x_{1},\ldots,x_{m},y))=\mu^{(\omega)}\left(\varphi(x_{1},\ldots,x_{m},x_{1}\cdot\ldots\cdot x_{k})\right).

Then λ|𝐱=μ(ω)\lambda|_{\mathbf{x}}=\mu^{(\omega)} and λ|(x1,,xk)y=λk\lambda|_{(x_{1},\ldots,x_{k})y}=\lambda_{k} from Lemma 3.44, so by Lemma 3.44 we then have |λ(φ(y,xi))μ(2)(φ(y,x))|>ε0|\lambda(\varphi(y,x_{i}))-\mu^{(2)}(\varphi(y,x))|>\varepsilon_{0} for all i=1,,ki=1,\ldots,k — contradicting the choice of nn. ∎

3.9. Support transitivity of idempotent measures

A tempting strategy for generalizing the arguments in Sections 2.72.10 with ranks from idempotent types to idempotent measures in TT is to apply (a continuous logic version of) the proof for types in the randomization TRT^{R} of TT, assuming that the randomization preserves the corresponding property. E.g, stability is preserved [BYK09], and (real) rosiness is known to be preserved in some special cases [AGK19] (e.g. when TT is oo-minimal). We note that simplicity of TT is not preserved, still one gets that TRT^{R} is NSOP1 assuming that TT is simple [BYCR24]. When attempting to implement this strategy, one arrives at the following natural condition connecting the behavior of measures and types in their support:

Definition 3.46.

Assume that GG is a type-definable group and μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}). We say that μ\mu is support transitive if μp=μ\mu\ast p=\mu for every pS(μ)p\in S(\mu).

Remark 3.47.
  1. (1)

    If pSG(𝒰)p\in S_{G}(\mathcal{U}) is a generically stable idempotent type, then it is obviously support transitive (viewed as a Keisler measure).

  2. (2)

    If μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is generically transitive, then it is support transitive.

    Indeed, note that if pS(μ)p\in S(\mu), θ(x,b)x(𝒰)\theta(x,b)\in\mathcal{L}_{x}(\mathcal{U}), μ\mu is MM-invariant for a small model M𝒰M\prec\mathcal{U}, and cp|Mbc\models p|_{Mb}, then

    (μp)(θ(x,b))=(μxpy)(θ(xy,b))=μ(θ(xc,b))=μ(θ(x,b)),(\mu*p)(\theta(x,b))=(\mu_{x}\otimes p_{y})(\theta(x\cdot y,b))=\mu(\theta(x\cdot c,b))=\mu(\theta(x,b)),

    where the last equality follows as pStab(μ)p\vdash\operatorname{Stab}(\mu) by assumption, and Stab(μ)\operatorname{Stab}(\mu) is MM-type-definable by Fact 3.34.

Thus, we view the following as an intermediate (and trivial in the case of types) version of our main Problem 3.41:

Problem 3.48.

Assume that GG is a type-definable group and μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is fim and idempotent. Is μ\mu support transitive?

The following example (based on [CG23, Example 4.5]) illustrates that the fim assumption in Problem 3.48 cannot be relaxed to either definable or Borel-definable and finitely satisfiable, even in abelian NIP groups:

Example 3.49.

Consider M:=(,<,+)M:=(\mathbb{R},<,+), M𝒰M\prec\mathcal{\mathcal{U}}, G(𝒰)=𝒰G(\mathcal{U})=\mathcal{U} and μ:=12δp+12δp\mu:=\frac{1}{2}\delta_{p_{-\infty}}+\frac{1}{2}\delta_{p_{\infty}} and ν:=12δp0+12δp0+\nu:=\frac{1}{2}\delta_{p_{0^{-}}}+\frac{1}{2}\delta_{p_{0^{+}}}, where p,p,p0+,p0p_{\infty},p_{-\infty},p_{0^{+}},p_{0^{-}} are the unique complete 11-types satisfying:

  • p{x>a:a}{x<b:b𝒰,b>}p_{\infty}\supseteq\{x>a:a\in\mathbb{R}\}\cup\{x<b:b\in\mathcal{U},b>\mathbb{R}\},

  • p{x<a:a}{x>b:b𝒰,b<}p_{-\infty}\supseteq\{x<a:a\in\mathbb{R}\}\cup\{x>b:b\in\mathcal{U},b<\mathbb{R}\},

  • p0+{x<a:a𝒰,a>0}{x>0}p_{0^{+}}\supseteq\{x<a:a\in\mathcal{U},a>0\}\cup\{x>0\},

  • p0{x<0}{x>b:b𝒰,b<0}p_{0^{-}}\supseteq\{x<0\}\cup\{x>b:b\in\mathcal{U},b<0\}.

Then μ\mu is finitely satisfiable in MM (hence also Borel-invariant over MM by NIP) and ν\nu is definable over MM, but neither is fim. The following are easy to verify directly:

  1. (1)

    μp=p\mu*p_{\infty}=p_{\infty}, μp=p\mu*p_{-\infty}=p_{-\infty} and μμ=μ\mu*\mu=\mu — hence μ\mu is idempotent, finitely satisfiable in MM, but not support-invariant;

  2. (2)

    likewise, νp0+=p0+\nu*p_{0^{+}}=p_{0^{+}}, νp0=p0\nu*p_{0^{-}}=p_{0^{-}}, and νν=ν\nu*\nu=\nu — hence ν\nu is idempotent, definable over MM, but not support invariant.

Support transitivity is closely related to the algebraic properties of the semigroup induced by \ast on the support of an idempotent measure, studied in [CG22, Section 4].

Fact 3.50.

[CG22, Corollary 4.4] Assume that μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is fim and idempotent. Then (S(μ),)(S(\mu),*) is a compact Hausdorff semigroup which is left-continuous, i.e. the map pS(μ)pqS(μ)p\in S(\mu)\mapsto p\ast q\in S(\mu) is continuous for each fixed qS(μ)q\in S(\mu).

Proposition 3.51.

Assume that μ𝔐G(𝒰)\mu\in\mathfrak{M}_{G}(\mathcal{U}) is fim and idempotent. Then the following are equivalent:

  1. (1)

    μ\mu is support transitive, i.e. μp=μ\mu\ast p=\mu for all pS(μ)p\in S(\mu);

  2. (2)

    for any p,qS(μ)p,q\in S(\mu) there exists rS(μ)r\in S(\mu) such that rq=pr\ast q=p;

  3. (3)

    Iμ=S(μ)I_{\mu}=S(\mu), where IμI_{\mu} is a minimal (closed) left ideal of (S(μ),)\left(S(\mu),\ast\right).

Proof.

(2) \Leftrightarrow (3). By [CG22, Remark 4.17].

(3) \Rightarrow (1). By [CG22, Corollary 4.16].

(1) \Rightarrow (2). Let p,qS(μ)p,q\in S(\mu) be given. By Fact 3.50, the map fp:S(μ)S(μ)f_{p}:S(\mu)\to S(\mu) defined via fp(s):=spf_{p}(s):=s\ast p is continuous. We will show that it has a dense image. Then by compactness of S(μ)S(\mu) and continuity of fpf_{p}, the image of fpf_{p} is also closed, hence fpf_{p} is surjective — proving the claim.

Indeed, fix some formula θ(x,c)x(𝒰)\theta(x,c)\in\mathcal{L}_{x}(\mathcal{U}) such that μ(θ(x,c))>0\mu(\theta(x,c))>0 and choose a small M𝒰M\prec\mathcal{U} such that μ\mu is MM-invariant. By (1), (μp)(θ(x,c))=μ(θ(x,c))>0(\mu\ast p)(\theta(x,c))=\mu(\theta(x,c))>0. Let bp|Mcb\models p|_{Mc}, then μ(θ(xb,c))>0\mu(\theta(x\cdot b,c))>0. Hence there exists some rS(μ)r\in S(\mu) such that θ(xb,c)r\theta(x\cdot b,c)\in r. But then by definition θ(x,c)rp\theta(x,c)\in r\ast p, hence fp(r)[θ(x,c)]f_{p}(r)\in[\theta(x,c)]. As θ(x,c)\theta(x,c) was arbitrary, this shows that the image of fpf_{p} is dense. ∎

We know that this property holds in specific examples like the circle group (e.g., see [CG22, Example 4.2]).

3.10. Idempotent measures in stable theories, revisited

It is shown in [CG22, Theorem 5.8] that every idempotent measure on a type-definable group in a stable theory is generically transitive. The proof consists of two ingredients: an analysis of the convolution semigroup on the support of an idempotent Keisler measure, and an application of a variant of Hrushovski’s group chunk theorem for partial types due to Newelski [New91].

In this section we provide an alternative argument, implementing the strategy outlined at the beginning of Section 3.9 of working in the randomization. This replaces the use of Newelski’s theorem by a direct generalization of the proof for types in stable theories from Section 2.7, and the only fact about the supports of idempotent measures that we will need is that they are support transitive.

To simplify the notation, in this section we will assume that TT is an \mathcal{L}-theory expanding a group. We first recall the basic results about local ranks in continuous logic, from [BY10, BYU10]. The following facts are proved under more general hypothesis in Sections 7 and 8 of [BYU10].

Fact 3.52.

Suppose that TT is a continuous stable theory. Let M𝒰TM\prec\mathcal{U}\models T.

  1. (1)

    For any pSx(M)p\in S_{x}(M) there exists a unique MM-definable extension pSx(𝒰)p^{\prime}\in S_{x}(\mathcal{U}).

  2. (2)

    For every ε>0\varepsilon>0 and every partitioned \mathcal{L}-formula φ(x,y)\varphi(x,y), there exists a rank function CBφ,ε\operatorname{CB}_{\varphi,\varepsilon}, which we call the ε\varepsilon-Cantor-Bendixson rank. More specifically, for any subset A𝒰A\subseteq\mathcal{U} and pSφ(A)p\in S_{\varphi}(A), CBφ,ε(p)CB_{\varphi,\varepsilon}(p) is an ordinal.

  3. (3)

    For any rSφ(M)r\in S_{\varphi}(M) and sSφ(𝒰)s\in S_{\varphi}(\mathcal{U}) such that srs\supseteq r, ss is the unique definable extension of rr if and only if for every ε>0\varepsilon>0, CBφ,ε(r)=CBφ,ε(s)\operatorname{CB}_{\varphi,\varepsilon}(r)=\operatorname{CB}_{\varphi,\varepsilon}(s).

The following proposition is a standard exercise from the previous fact.

Proposition 3.53.

Let TT be a continuous stable theory expanding a group. Let 𝒰\mathcal{U} be a monster model of TT and M𝒰M\prec\mathcal{U} a small submodel. For any partitioned \mathcal{L}-formula φ(x,y)\varphi(x,y) we let Δφ:=φ(x;y,z)=φ(xz,y)\Delta_{\varphi}:=\varphi(x;y,z)=\varphi(x\cdot z,y). Then:

  1. (1)

    for any rSx(𝒰)r\in S_{x}(\mathcal{U}) and g𝒰g\in\mathcal{U}, CBΔφ,ε(r|Δφ)=CBΔφ,ε(rg|Δφ)\operatorname{CB}_{\Delta_{\varphi},\varepsilon}(r|_{\Delta_{\varphi}})=\operatorname{CB}_{\Delta_{\varphi},\varepsilon}(r\cdot g|_{\Delta_{\varphi}});

  2. (2)

    for any pSx(M)p\in S_{x}(M) and qSx(𝒰)q\in S_{x}(\mathcal{U}) such that qpq\supset p, we have that qq is the unique definable extension of pp if and only if for every partitioned \mathcal{L}-formula φ(x,y)\varphi(x,y) and for all ε>0\varepsilon>0, we have CBΔφ,ε(q|Δφ)=CBΔφ,ε(p|Δφ)\operatorname{CB}_{\Delta_{\varphi},\varepsilon}(q|_{\Delta_{\varphi}})=\operatorname{CB}_{\Delta_{\varphi},\varepsilon}(p|_{\Delta_{\varphi}}).

Proof.
  1. (1)

    Given g𝒰g\in\mathcal{U}, the map mg:SΔφ(𝒰)SΔφ(𝒰)m_{g}:S_{\Delta_{\varphi}}(\mathcal{U})\to S_{\Delta_{\varphi}}(\mathcal{U}) defined via φ(xa,b)mg(r)=φ(xga,b)r\varphi(x\cdot a,b)^{m_{g}(r)}=\varphi(x\cdot g\cdot a,b)^{r} is a bijective isometry. In other words, it is an automorphism of SΔφ(𝒰)S_{\Delta_{\varphi}}(\mathcal{U}) as a topometric space, and computing the rank is unaffected.

  2. (2)

    Follows directly from (3) of Fact 3.52. ∎

We refer to Section 3.4 for notation regarding Keisler randomizations.

Fact 3.54.

[BYK09, Theorem 5.14] If TT is stable, then its Keisler randomization TRT^{R} is stable.

Proposition 3.55.

Suppose that TT is stable, μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) is idempotent and support transitive. Then μ\mu is generically transitive.

Proof.

Let 𝒱𝒰\mathcal{V}\succ\mathcal{U} be a bigger monster model of TT. Fix an atomless probability algebra (Ω,,)(\Omega,\mathcal{B},\mathbb{P}) and consider the randomizations 𝒰Ω𝒱Ω𝒞\mathcal{U}^{\Omega}\prec\mathcal{V}^{\Omega}\prec\mathcal{C}, where 𝒞\mathcal{C} is a monster model of TRT^{R}. Given μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) (note that μ\mu is definably by stability), we let rμ𝒰SxR(𝒞)r_{\mu}^{\mathcal{U}}\in S_{x}^{R}(\mathcal{C}) be as defined in Fact 3.17. Similarly, given μ𝔐x(𝒱)\mu\in\mathfrak{M}_{x}(\mathcal{V}), we let rμ𝒱SxR(𝒞)r_{\mu}^{\mathcal{V}}\in S_{x}^{R}(\mathcal{C}) be as defined in Fact 3.17, but with respect to 𝒱\mathcal{V} in place of 𝒰\mathcal{U}.

Let now μ𝔐x(𝒰)\mu\in\mathfrak{M}_{x}(\mathcal{U}) be idempotent and support transitive. Since TT is stable, there is some small model M𝒰M\prec\mathcal{U} such that μ\mu is MM-definable. Let μ𝔐x(𝒱)\mu^{\prime}\in\mathfrak{M}_{x}(\mathcal{V}) be the unique MM-definable extension of μ\mu. To show that μ\mu is generically transitive, it suffices to prove that for every pS(μ)p\in S(\mu) and a𝒱a\in\mathcal{V} such that apa\models p, we have that μ=μa\mu^{\prime}=\mu^{\prime}\cdot a. We let 𝐩:=rμ𝒰|𝒰Ω\mathbf{p}:=r_{\mu}^{\mathcal{U}}|_{\mathcal{U}^{\Omega}}. By construction, rμ𝒱𝐩r_{\mu^{\prime}}^{\mathcal{V}}\supseteq\mathbf{p} and rμ𝒱r_{\mu^{\prime}}^{\mathcal{V}} is MΩM^{\Omega}-definable. Since TRT^{R} is stable (Fact 3.54), rμ𝒱r_{\mu^{\prime}}^{\mathcal{V}} is the unique global definable extension of 𝐩\mathbf{p}.

We claim that then rμa𝒱𝐩r_{\mu^{\prime}\cdot a}^{\mathcal{V}}\supseteq\mathbf{p}. Indeed, let φ(x,y)\varphi(x,y) be an \mathcal{L}-formula and h𝒰0Ωh\in\mathcal{U}^{\Omega}_{0}. If 𝒜\mathcal{A} is a partition of Ω\Omega for hh, using that μ\mu is support transitive we have:

(𝔼[φ(x,h)])rμa𝒱=A𝒜(A)(μa)(φ(x,h|A))=A𝒜(A)(μp)(φ(x,h|A))\displaystyle(\mathbb{E}[\varphi(x,h)])^{r_{\mu^{\prime}\cdot a}^{\mathcal{V}}}=\sum_{A\in\mathcal{A}}\mathbb{P}(A)(\mu^{\prime}\cdot a)(\varphi(x,h|_{A}))=\sum_{A\in\mathcal{A}}\mathbb{P}(A)(\mu*p)(\varphi(x,h|_{A}))
=A𝒜(A)μ(φ(x,h|A))=A𝒜(A)μ(φ(x,h|A))=(𝔼[φ(x,h)])rμ𝒱.\displaystyle=\sum_{A\in\mathcal{A}}\mathbb{P}(A)\mu(\varphi(x,h|_{A}))=\sum_{A\in\mathcal{A}}\mathbb{P}(A)\mu^{\prime}(\varphi(x,h|_{A}))=(\mathbb{E}[\varphi(x,h)])^{r_{\mu^{\prime}}^{\mathcal{V}}}.

Likewise, it is straightforward to check that rμa𝒱=rμ𝒱far_{\mu^{\prime}\cdot a}^{\mathcal{V}}=r_{\mu^{\prime}}^{\mathcal{V}}\cdot f_{a}, where fa𝒰0Ωf_{a}\in\mathcal{U}_{0}^{\Omega} is the constant random variable (i.e., fa:Ω𝒰f_{a}:\Omega\to\mathcal{U} via fa(t)=af_{a}(t)=a for all tΩt\in\Omega) and \cdot is the randomization of the multiplication of the group in TT.

Since local rank in the stable theory TRT^{R} is translation invariant (Proposition 3.53), we conclude that rμa𝒱r_{\mu^{\prime}\cdot a}^{\mathcal{V}} is the unique definable extension of 𝐩\mathbf{p}. This implies that rμ𝒱=rμa𝒱r_{\mu^{\prime}}^{\mathcal{V}}=r_{\mu^{\prime}\cdot a}^{\mathcal{V}} and in turn, μ=μa\mu^{\prime}=\mu^{\prime}\cdot a. This completes the proof. ∎

Remark 3.56.

We expect that this approach could be adapted for groups definable in oo-minimal structures (as their randomizations are known to be real rosy [AGK19]), by developing a stratified local thorn rank in continuous logic and generalizing the proof for types in rosy (discrete) first order theories from Section 2.10. When TT is a simple theory, the randomization TRT^{R} is NSOP1 (but not necessarily simple) by [BYCR24]. A local rank for NSOP1 theories is proposed in [CKR23, Section 5], but a workable stratified rank is lacking at the moment. We do not pursue these directions here.

4. Topological dynamics of 𝔐xfs(𝒢,G)\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G) and Sxfs(𝒢,G)S_{x}^{\operatorname{fs}}(\mathcal{G},G) in NIP groups

In this section, we will use slightly different notation from the rest of the paper, in order to preserve continuity with the earlier work and setup in [CG22, CG23]. We let GG be an expansion of a group, and 𝒢G\mathcal{G}\succ G a monster model. Throughout this section, we assume that T:=Th(G)T:=\operatorname{{Th}}(G) has NIP.

It was demonstrated in [CG22, Proposition 6.4] that then the spaces of global Aut(𝒢/G)\operatorname{Aut}(\mathcal{G}/G)-invariant Keisler measures, and Keisler measures which are finitely satisfiable in GG (denoted 𝔐xinv(𝒢,G)\mathfrak{M}_{x}^{\operatorname{inv}}(\mathcal{G},G) and 𝔐xfs(𝒢,G)\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G), respectively) form left-continuous compact Hausdorff semigroups with respect to definable convolution (Definition 3.35). Note that (𝔐xfs(𝒢,G),)(\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G),*) is a submonoid of (𝔐xinv(𝒢,G),)(\mathfrak{M}_{x}^{\operatorname{inv}}(\mathcal{G},G),*). By Sxfs(𝒢,G)S_{x}^{\operatorname{fs}}(\mathcal{G},G) we denote the submonoid of (𝔐xfs(𝒢,G),)(\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G),*) consisting of all global types finitely satisfiable in GG (viewed as {0,1}\{0,1\}-measures).

In [CG23, Theorem 6.11], the first two authors described a minimal left ideal of (𝔐xfs(𝒢,G),)(\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G),*) [and (𝔐xinv(𝒢,G),)(\mathfrak{M}_{x}^{\operatorname{inv}}(\mathcal{G},G),*)] in terms of the Haar measure on an ideal (or Ellis) group of (Sxfs(𝒢,G),)(S_{x}^{\operatorname{fs}}(\mathcal{G},G),*) [resp. (Sxinv(𝒢,G),)(S_{x}^{\operatorname{inv}}(\mathcal{G},G),*)]. However, this required a rather specific assumption that this ideal group is a compact topological group with the topology induced from (Sxfs(𝒢,G),)(S_{x}^{\operatorname{fs}}(\mathcal{G},G),*) [resp. (Sxinv(𝒢,G),)(S_{x}^{\operatorname{inv}}(\mathcal{G},G),*)]. In this section, we obtain the same description in the case of (𝔐xfs(𝒢,G),)(\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G),*), but under a more natural (from the point of view of topological dynamics) assumption that the so-called τ\tau-topology on some (equivalently, every) ideal group is Hausdorff (equivalently, the ideal group with the τ\tau-topology is a compact topological group). In fact, the revised Newelski’s conjecture formulated by Anand Pillay and the third author in [KP23, Conjecture 5.3] predicts that the τ\tau-topology is always Hausdorff under NIP. In Section 5, we confirm this conjecture in the case when GG is countable, which is an important result by its own rights. In particular, in the case when GG is countable, our description of a minimal left ideal of (𝔐xfs(𝒢,G),)(\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G),*) does not require any assumption on the ideal group.

As discussed in the introduction, the τ\tau-topology plays an essential role in many important structural results in abstract topological dynamics, including the recent theorem of Glasner on the structure of tame, metrizable, minimal flows [Gla18]. In fact, our proof of the revised Newelski’s conjecture for countable GG will be deduced using this theorem of Glasner.

The reason why our proof of the revised Newelski’s conjecture requires the countability of GG assumption is to guarantee that certain flows of types are metrizable in order to be able to apply the aforementioned theorem of Glasner. The reason why we focus only on (𝔐xfs(𝒢,G),)(\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G),*) and (Sxfs(𝒢,G),)(S_{x}^{\operatorname{fs}}(\mathcal{G},G),*) (and not on (𝔐xinv(𝒢,G),)(\mathfrak{M}_{x}^{\operatorname{inv}}(\mathcal{G},G),*) and (Sxinv(𝒢,G),)(S_{x}^{\operatorname{inv}}(\mathcal{G},G),*)) is that (Sxfs(𝒢,G),)(S_{x}^{\operatorname{fs}}(\mathcal{G},G),*) is isomorphic to the Ellis semigroup of the GG-flow Sxfs(𝒢,G)S_{x}^{\operatorname{fs}}(\mathcal{G},G) and so we have the τ\tau-topology on the ideal group of Sxfs(𝒢,G)S_{x}^{\operatorname{fs}}(\mathcal{G},G) at our disposal. For the revised Newelski’s conjecture we will also use a well-known general principle that NIP implies tameness for various flows of types [CS18, Iba16, KR20].

4.1. Preliminaries from topological dynamics

Definition 4.1.

A GG-flow is a pair (G,X)(G,X), where GG is an abstract group acting (on the left) by homeomorphisms on a compact Hausdorff space XX.

Definition 4.2.

If (G,X)(G,X) is a flow, then its Ellis semigroup, denoted by E(G,X)E(G,X) or just E(X)E(X), is the (pointwise) closure in XXX^{X} of the set of functions πg:xgx\pi_{g}\colon x\mapsto g\cdot x for gGg\in G.

Fact 4.3.

(see e.g. [Aus88]) If (G,X)(G,X) is a flow, then E(X)E(X) is a compact left topological semigroup (i.e. it is a semigroup with the composition as its semigroup operation, and the composition is continuous on the left, i.e. for any fE(X)f\in E(X) the map f-\circ f is continuous). It is also a GG-flow with gf:=πgfg\cdot f:=\pi_{g}\circ f.

The next fact is folklore. Thanks to this fact E(E(X))E(E(X)) is always identified with E(X)E(X).

Fact 4.4.

The function Φ:E(X)E(E(X))\Phi\colon E(X)\to E(E(X)) given by Φ(η):=lη\Phi(\eta):=l_{\eta}, where lη:E(X)E(X)l_{\eta}\colon E(X)\to E(X) is defined by lη(τ):=ητl_{\eta}(\tau):=\eta\circ\tau, is an isomorphism of semigroups and GG-flows.

The following is a fundamental theorem of Ellis on the basic structure of Ellis semigroups (see e.g. [Ell69, Corollary 2.10 and Propositions 3.5 and 3.6] or Proposition 2.3 of [Gla76, Section I.2]). We will use it freely without an explicit reference.

Fact 4.5 (Ellis’ Theorem).

Suppose SS is a compact Hausdorff left topological semigroup (e.g. the enveloping semigroup of a flow). Then SS has a minimal left ideal \mathcal{M}. Furthermore, for any such ideal \mathcal{M}:

  1. (1)

    \mathcal{M} is closed;

  2. (2)

    for any element aa\in\mathcal{M} and idempotent uu\in\mathcal{M} we have au=aau=a, and =Sa=a\mathcal{M}=Sa=\mathcal{M}a;

  3. (3)

    =uu\mathcal{M}=\bigsqcup_{u}u\mathcal{M}, where uu ranges over all idempotents in \mathcal{M}; in particular, \mathcal{M} contains an idempotent;

  4. (4)

    for any idempotent uu\in\mathcal{M}, the set uu\mathcal{M} is a subgroup of SS with the neutral element uu.

Moreover, all the groups uu\mathcal{M} (where \mathcal{M} ranges over all minimal left ideals and uu over all idempotents in \mathcal{M}) are isomorphic. In the model theory literature, the isomorphism type of all these groups (or any of these groups) is called the ideal (or Ellis) group of SS; if S=E(G,X)S=E(G,X), we call this group the ideal (or Ellis) group of the flow (G,X)(G,X).

We will use the following fact, which gives us an explicit isomorphism from Fact 4.5 between any two ideal groups in a given minimal left ideal. The context is as in Fact 4.5.

Fact 4.6.

If uu and vv are idempotents in \mathcal{M}, then pupp\mapsto up defines an isomorphism lu:vul_{u}\colon v\mathcal{M}\to u\mathcal{M}.

Proof.

The map lul_{u} is a homomorphism, because lu(pq)=u(pq)=u(pu)q=(up)(uq)=lu(p)lu(q)l_{u}(pq)=u(pq)=u(pu)q=(up)(uq)=l_{u}(p)l_{u}(q). In the same way, lv:uvl_{v}\colon u\mathcal{M}\to v\mathcal{M} given by lv(p):=vpl_{v}(p):=vp is a homomorphism. And it is clear that lvl_{v} is the inverse of lul_{u}. ∎

We will also need the following observation which was Lemma 3.5 in the first arXiv version of [KLM22] (the section with this results was removed in the published version).

Lemma 4.7.

Let SS be a compact left topological semigroup, \mathcal{M} a minimal left ideal of SS, and uu\in\mathcal{M} an idempotent. Then the closure u¯\overline{u\mathcal{M}} of uu\mathcal{M} is a (disjoint) union of ideal groups. In particular, u¯\overline{u\mathcal{M}} is a subsemigroup of SS.

Proof.

Note that for every η\eta\in\mathcal{M}, η\eta\mathcal{M} is an ideal group. Namely, ηv\eta\in v\mathcal{M} for some idempotent vv\in\mathcal{M}. Thus, ηv\eta\mathcal{M}\subseteq v\mathcal{M}, but also vηv\mathcal{M}\subseteq\eta\mathcal{M}, because v=ηη1v=\eta\eta^{-1}, where η1\eta^{-1} is the inverse of η\eta in the ideal group vv\mathcal{M}.

Let now η0u¯\eta_{0}\in\overline{u\mathcal{M}}\subseteq\mathcal{M} (as \mathcal{M} is closed). By the first paragraph, it suffices to prove that η0u¯\eta_{0}\mathcal{M}\subseteq\overline{u\mathcal{M}}. Since η0u=η0\eta_{0}u=\eta_{0}, we have η0u=η0\eta_{0}u\mathcal{M}=\eta_{0}\mathcal{M}. Take any ηη0\eta\in\eta_{0}\mathcal{M}. Then η=η0η\eta=\eta_{0}\eta^{\prime} for some ηu\eta^{\prime}\in u\mathcal{M}. Since η0u¯\eta_{0}\in\overline{u\mathcal{M}}, we have that η0\eta_{0} is the limit point of a net (η0i)iu(\eta_{0}^{i})_{i}\subseteq u\mathcal{M}. By left continuity, η=η0η=limi(η0iη)\eta=\eta_{0}\eta^{\prime}=\lim_{i}\left(\eta_{0}^{i}\eta^{\prime}\right), so ηu¯\eta\in\overline{u\mathcal{M}} as η0iηu\eta_{0}^{i}\eta^{\prime}\in u\mathcal{M} for all ii’s. ∎

Most of the statements in the next fact are contained in [Gla76, Section IX.1]. There, the author considers the special case of X=βGX=\beta G and defines \circ in a slightly different (but equivalent) way. However, as pointed out in [KP17, Section 2] and [KPR18, Section 1.1], many of the proofs from [Gla76, Section IX.1] go through in the general context. A very nice exposition of this material (with all the proofs) in the general context can be found in Appendix A of [Rze18].

Fact 4.8 (The τ\tau-topology on the ideal group in an Ellis semigroup).

Consider the Ellis semigroup E(X)E(X) of a flow (G,X)(G,X), let \mathcal{M} be a minimal left ideal of E(X)E(X) and uu\in\mathcal{M} an idempotent.

  1. (1)

    For each aE(X)a\in E(X), BE(X)B\subseteq E(X), we write aBa\circ B for the set of all limits of nets (gibi)i(g_{i}b_{i})_{i}, where giGg_{i}\in G are such that limigi=a\lim_{i}g_{i}=a and biBb_{i}\in B.

  2. (2)

    The formula clτ(A):=(u)(uA)\operatorname{cl}_{\tau}(A):=(u\mathcal{M})\cap(u\circ A) defines a closure operator on uu\mathcal{M}. It can also be (equivalently) defined as clτ(A)=u(uA)\operatorname{cl}_{\tau}(A)=u(u\circ A). We call the topology on uu\mathcal{M} induced by this operator the τ\tau-topology.

  3. (3)

    If (fi)i(f_{i})_{i} (a net in uu\mathcal{M}) converges to fu¯f\in\overline{u\mathcal{M}} (the closure of uu\mathcal{M} in E(X)E(X)), then (fi)i(f_{i})_{i} converges to ufuf in the τ\tau-topology.

  4. (4)

    The τ\tau-topology on uu\mathcal{M} coarsens the subspace topology inherited from E(X)E(X).

  5. (5)

    uu\mathcal{M} with the τ\tau-topology is a quasi-compact, T1T_{1} semitopological group (that is, the group operation is separately continuous) in which the inversion is continuous.

  6. (6)

    All the groups uu\mathcal{M} (where \mathcal{M} ranges over all minimal left ideals of E(X)E(X) and uu over all idempotents in \mathcal{M}) equipped with the τ\tau-topology are isomorphic as semitopological groups. In particular, the map from Fact 4.6 is a topological isomorphism.

By the Ellis joint continuity theorem [Ell57] and Fact 4.8(5), we get the following

Corollary 4.9.

If the τ\tau-topology on uu\mathcal{M} is Hausdorff, then uu\mathcal{M} is a compact topological group.

The following result is Lemma 3.1 in [KPR18].

Fact 4.10.

Let (G,X)(G,X) be a flow. Let \mathcal{M} be a minimal left ideal of E(X)E(X) and uu\in\mathcal{M} an idempotent. Then the function f:u¯uf\colon\overline{u\mathcal{M}}\to u\mathcal{M} (where u¯\overline{u\mathcal{M}} is the closure of uu\mathcal{M} in the topology of E(X)E(X)) defined by the formula f(η):=uηf(\eta):=u\eta has the property that for any continuous function h:uXh\colon u\mathcal{M}\to X, where XX is a regular topological space and uu\mathcal{M} is equipped with the τ\tau-topology, the composition hf:u¯Xh\circ f\colon\overline{u\mathcal{M}}\to X is continuous, where u¯\overline{u\mathcal{M}} is equipped with subspace topology from E(X)E(X). In particular, if uu\mathcal{M} is Hausdorff with the τ\tau-topology, then ff is continuous.

In the model-theoretic context of the GG-flow Sxfs(𝒢,G)S_{x}^{\operatorname{fs}}(\mathcal{G},G) (with the action of GG by left translations), the following fact is folklore (see e.g. [New12, Pil13]).

Fact 4.11.

The function Φ:Sxfs(𝒢,G)E(Sxfs(𝒢,G))\Phi\colon S_{x}^{\operatorname{fs}}(\mathcal{G},G)\to E(S_{x}^{\operatorname{fs}}(\mathcal{G},G)) given by Φ(p):=lp\Phi(p):=l_{p}, where lp:Sxfs(𝒢,G)Sxfs(𝒢,G)l_{p}\colon S_{x}^{\operatorname{fs}}(\mathcal{G},G)\to S_{x}^{\operatorname{fs}}(\mathcal{G},G) is defined by lp(q):=pql_{p}(q):=p*q, is an isomorphism of semigroups and GG-flows.

Thanks to this fact, Fact 4.8 can be applied directly to Sxfs(𝒢,G)S_{x}^{\operatorname{fs}}(\mathcal{G},G) in place of E(Sxfs(𝒢,G))E(S_{x}^{\operatorname{fs}}(\mathcal{G},G)), which we do without further explanations.

Note, however, that Fact 4.11 does not hold for Sxinv(𝒢,G)S_{x}^{\operatorname{inv}}(\mathcal{G},G) in place of Sxfs(𝒢,G)S_{x}^{\operatorname{fs}}(\mathcal{G},G), because the GG-orbit of tp(e/𝒢)\operatorname{tp}(e/\mathcal{G}) need not be dense in Sxinv(𝒢,G)S_{x}^{\operatorname{inv}}(\mathcal{G},G).

4.2. Minimal left ideal of 𝔐xfs(𝒢,G)\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G)

Recall that GG is an expansion of a group. Fix 𝒢G\mathcal{G}\succ G which is |G|+|G|^{+}-saturated. Recall that in this section we assume that T=Th(G)T=\operatorname{{Th}}(G) is NIP.

Definition 4.12.

For a formula φ(x)(𝒢)\varphi(x)\in\mathcal{L}(\mathcal{G}) and n>0n>0, define a new formula

Altn(x0,,xn1;y):=i<n1¬(φ(xiy)φ(xi+1y)).\operatorname{{Alt}}_{n}(x_{0},\dots,x_{n-1};y):=\bigwedge_{i<n-1}\neg(\varphi(x_{i}y)\leftrightarrow\varphi(x_{i+1}y)).

In the next fact, x¯\bar{x} stands for (x0,,xn1)(x_{0},\dots,x_{n-1}). By the proof of [HP11, Proposition 2.6] (where, if φ(x)(𝒢)\varphi(x)\in\mathcal{L}(\mathcal{G}) is of the form ψ(x;c)\psi(x;c) for some ψ(x,y)()\psi(x,y)\in\mathcal{L}(\emptyset) and c𝒢yc\in\mathcal{G}^{y}, NN is chosen depending on ψ(x,y)\psi(x,y)), we have:

Fact 4.13.

Let pSxinv(𝒢,G)p\in S_{x}^{\operatorname{inv}}(\mathcal{G},G) and φ(x)(𝒢)\varphi(x)\in\mathcal{L}(\mathcal{G}) be any formula. Let S:={b𝒢:φ(xb)p}S:=\{b\in\mathcal{G}:\varphi(xb)\in p\}. Then, there exists a positive N<ωN<\omega such that S=n<NAnBn+1cS=\bigcup_{n<N}A_{n}\cap B_{n+1}^{c}, where c-^{c} denotes the complement of a set and

An:={b𝒢:(x¯)(p(n)|G(x¯)Altn(x¯;b)φ(xn1b))},Bn:={b𝒢:(x¯)(p(n)|G(x¯)Altn(x¯;b)¬φ(xn1b))}.\begin{array}[]{ll}A_{n}:=\{b\in\mathcal{G}:(\exists\bar{x})(p^{(n)}|_{G}(\bar{x})\wedge\operatorname{{Alt}}_{n}(\bar{x};b)\wedge\varphi(x_{n-1}b))\},\\ B_{n}:=\{b\in\mathcal{G}:(\exists\bar{x})(p^{(n)}|_{G}(\bar{x})\wedge\operatorname{{Alt}}_{n}(\bar{x};b)\wedge\neg\varphi(x_{n-1}b))\}.\end{array}

The following lemma is the key new step needed to adapt the arguments from [CG23, Section 6.2] to our general setting here. From now on, let \mathcal{M} be a minimal left ideal in (Sxfs(𝒢,G),)(S_{x}^{\operatorname{fs}}(\mathcal{G},G),*) and uu\in\mathcal{M} an idempotent. For p,qSxfs(𝒢,G)p,q\in S_{x}^{\operatorname{fs}}(\mathcal{G},G), we will typically write pqpq instead of pqp\ast q.

Lemma 4.14.

Let φ(x)(𝒢)\varphi(x)\in\mathcal{L}(\mathcal{G}) be any formula. Assume that the τ\tau-topology on the ideal group uu\mathcal{M} is Hausdorff. Then the subset [φ(x)]u[\varphi(x)]\cap u\mathcal{M} of uu\mathcal{M} is constructible, and so Borel (in the τ\tau-topology).

Proof.

By Fact 4.10 and the assumption that uu\mathcal{M} is Hausdorff, the function f:u¯uf\colon\overline{u\mathcal{M}}\to u\mathcal{M} (where u¯\overline{u\mathcal{M}} is the closure of uu\mathcal{M} in the topology of E(X)E(X)) given by f(η):=uηf(\eta):=u\eta is continuous. Note that ker(f):={pu¯:f(p)=u}\ker(f):=\{p\in\overline{u\mathcal{M}}:f(p)=u\} is a subsemigroup of u¯\overline{u\mathcal{M}} which coincides with the set 𝒥\mathcal{J} of all idempotents in u¯\overline{u\mathcal{M}}. This follows from Lemma 4.7 and the fact that for every v𝒥v\in\mathcal{J} the restriction f|v:vuf|_{v\mathcal{M}}\colon v\mathcal{M}\to u\mathcal{M} is a group isomorphism by Fact 4.6.

Put

S~:=f1[[φ(x)]]={pu¯:up[φ(x)]}.\widetilde{S}:=f^{-1}[[\varphi(x)]]=\{p\in\overline{u\mathcal{M}}:up\in[\varphi(x)]\}.

Let 𝔊𝒢{\mathfrak{G}}\succ\mathcal{G} be a monster model in which 𝒢\mathcal{G} is small. Let u¯Sxfs(𝔊,G)\bar{u}\in S_{x}^{\operatorname{fs}}({\mathfrak{G}},G) be the unique extension of uu to a type in Sx(𝔊)S_{x}({\mathfrak{G}}) which is finitely satisfiable in GG. Pick au¯a\models\bar{u} (in a yet bigger monster model). Put

S:={b𝔊:φ(ab)}={b𝔊:φ(xb)u¯}.S:=\{b\in{\mathfrak{G}}:\models\varphi(ab)\}=\{b\in{\mathfrak{G}}:\varphi(xb)\in\bar{u}\}.

Then S~={pu¯:φ(ab) for all/some bp(𝔊)}={tp(b/𝒢)u¯:bS}\widetilde{S}=\{p\in\overline{u\mathcal{M}}:\models\varphi(ab)\textrm{ for all/some }b\in p({\mathfrak{G}})\}=\{\operatorname{tp}(b/\mathcal{G})\in\overline{u\mathcal{M}}:b\in S\}.

Take NN, AnA_{n}, and BnB_{n} from Fact 4.13 applied for 𝔊{\mathfrak{G}} in place of 𝒢\mathcal{G} and for p=u¯p=\bar{u}. Then S=n<NAnBn+1cS=\bigcup_{n<N}A_{n}\cap B_{n+1}^{c}. Define:

A~n:={tp(b/𝒢):bAn and tp(b/𝒢)u¯},B~n:={tp(b/𝒢):bBn and tp(b/𝒢)u¯},A~n:=ker(f)A~n:={rs:rker(f),sA~n},B~n:=ker(f)B~n, and S~:=n<NA~nB~n+1c.\begin{array}[]{ll}\widetilde{A}_{n}:=\{\operatorname{tp}(b/\mathcal{G}):b\in A_{n}\textrm{ and }\operatorname{tp}(b/\mathcal{G})\in\overline{u\mathcal{M}}\},\\ \widetilde{B}_{n}:=\{\operatorname{tp}(b/\mathcal{G}):b\in B_{n}\textrm{ and }\operatorname{tp}(b/\mathcal{G})\in\overline{u\mathcal{M}}\},\\ \widetilde{A}_{n}^{\prime}:=\ker(f)*\widetilde{A}_{n}:=\left\{rs:r\in\ker(f),s\in\widetilde{A}_{n}\right\},\widetilde{B}_{n}^{\prime}:=\ker(f)*\widetilde{B}_{n},\textrm{ and }\\ \widetilde{S}^{\prime}:=\bigcup_{n<N}\widetilde{A}_{n}^{\prime}\cap\widetilde{B}_{n+1}^{\prime c}.\end{array}

Note that all the sets A~n,B~n,A~n,B~n,S~\widetilde{A}_{n},\widetilde{B}_{n},\widetilde{A}_{n}^{\prime},\widetilde{B}_{n}^{\prime},\widetilde{S}^{\prime} are contained in u¯\overline{u\mathcal{M}}, as u¯\overline{u\mathcal{M}} is a semigroup by Lemma 4.7.

Claim.
  1. (1)

    S~=n<NA~nB~n+1c\widetilde{S}=\bigcup_{n<N}\widetilde{A}_{n}\cap\widetilde{B}_{n+1}^{c}.

  2. (2)

    ker(f)S~=S~\ker(f)*\widetilde{S}=\widetilde{S}.

  3. (3)

    ker(f)A~n=A~n\ker(f)*\widetilde{A}_{n}^{\prime}=\widetilde{A}_{n}^{\prime}, ker(f)B~n=B~n\ker(f)*\widetilde{B}_{n}^{\prime}=\widetilde{B}_{n}^{\prime}, and ker(f)S~=S~\ker(f)*\widetilde{S}^{\prime}=\widetilde{S}^{\prime}.

  4. (4)

    A~nA~n\widetilde{A}_{n}\subseteq\widetilde{A}_{n}^{\prime} and B~nB~n\widetilde{B}_{n}\subseteq\widetilde{B}_{n}^{\prime}.

  5. (5)

    S~=S~\widetilde{S}=\widetilde{S}^{\prime}.

  6. (6)

    f[A~n]f[\widetilde{A}_{n}] and f[B~n]f[\widetilde{B}_{n}] are closed.

  7. (7)

    f[S~]=n<Nf[A~n]f[B~n+1]c=n<Nf[A~n]f[B~n+1]cf[\widetilde{S}^{\prime}]=\bigcup_{n<N}f[\widetilde{A}_{n}^{\prime}]\cap f[\widetilde{B}_{n+1}^{\prime}]^{c}=\bigcup_{n<N}f[\widetilde{A}_{n}]\cap f[\widetilde{B}_{n+1}]^{c} is constructible.

Proof.

(1) Let ρ:𝔊Sx(𝒢)\rho\colon{\mathfrak{G}}\to S_{x}(\mathcal{G}) be given by ρ(b):=tp(b/𝒢)\rho(b):=\operatorname{tp}(b/\mathcal{G}). Note that all of the sets An,BnA_{n},B_{n} are Aut(𝔊/𝒢)\operatorname{Aut}({\mathfrak{G}}/\mathcal{G})-invariant (more precisely, invariant over GG and the parameters of φ(x)\varphi(x) in 𝒢\mathcal{G}), hence they are unions of sets of realizations in 𝔊{\mathfrak{G}} of complete types over 𝒢\mathcal{G}. Then we have

ρ[S]=ρ[n<NAnBn+1c]=n<Nρ[An]ρ[Bn+1]c.\rho[S]=\rho\left[\bigcup_{n<N}A_{n}\cap B_{n+1}^{c}\right]=\bigcup_{n<N}\rho[A_{n}]\cap\rho[B_{n+1}]^{c}.

So S~=u¯ρ[S]=n<N(u¯ρ[An])(u¯ρ[Bn+1])c=n<NA~nB~n+1c\widetilde{S}=\overline{u\mathcal{M}}\cap\rho[S]=\bigcup_{n<N}(\overline{u\mathcal{M}}\cap\rho[A_{n}])\cap(\overline{u\mathcal{M}}\cap\rho[B_{n+1}])^{c}=\bigcup_{n<N}\widetilde{A}_{n}\cap\widetilde{B}_{n+1}^{c}.

(2) ()(\subseteq) Take pker(f)p\in\ker(f) and sS~s\in\widetilde{S}. Then f(ps)=u(ps)=(up)s=us=f(s)[φ(x)]f(ps)=u(ps)=(up)s=us=f(s)\in[\varphi(x)], which by definition implies psS~ps\in\widetilde{S}.

()(\supseteq) Take pS~p\in\widetilde{S}. Then, by Lemma 4.7, pvp\in v\mathcal{M} for some v𝒥v\in\mathcal{J} (where 𝒥\mathcal{J} is the set of all idempotents in u¯\overline{u\mathcal{M}}). Then p=vpp=vp and vker(f)v\in\ker(f), so pker(f)S~p\in\ker(f)*\widetilde{S}.

(3) The first two equalities follow from the definition of AnA_{n}^{\prime} and BnB_{n}^{\prime} and the fact that ker(f)=𝒥\ker(f)=\mathcal{J} satisfies 𝒥𝒥=𝒥\mathcal{J}*\mathcal{J}=\mathcal{J} (as uv=uuv=u for u,v𝒥u,v\in\mathcal{J}). To show ()(\subseteq) in the third equality, take any pA~nB~n+1cp\in\widetilde{A}_{n}^{\prime}\cap\widetilde{B}_{n+1}^{\prime c} and vker(f)=𝒥v\in\ker(f)=\mathcal{J}. Then vpA~nvp\in\widetilde{A}_{n}^{\prime} by the first equality. Moreover, vpB~n+1cvp\in\widetilde{B}_{n+1}^{\prime c}, as otherwise vpB~n+1vp\in\widetilde{B}_{n+1}^{\prime}, so p=vpvpB~n+1p=v_{p}vp\in\widetilde{B}_{n+1}^{\prime} by the second equality, where vp𝒥v_{p}\in\mathcal{J} is such that pvpp\in v_{p}\mathcal{M}, a contradiction. To see ()(\supseteq), take any pS~p\in\widetilde{S}^{\prime}. We have that pvp\in v\mathcal{M} for some v𝒥=ker(f)v\in\mathcal{J}=\ker(f). So p=vpker(f)S~p=vp\in\ker(f)*\widetilde{S}^{\prime}.

(4) follows as on the last line of (3).

(5) ()(\supseteq) Take gS~g\in\widetilde{S}^{\prime}. Then gA~nB~n+1cg\in\widetilde{A}_{n}^{\prime}\cap\widetilde{B}_{n+1}^{\prime c} for some n<Nn<N. Hence, gker(f)hg\in\ker(f)*h for some hA~nh\in\widetilde{A}_{n}. Since gB~n+1cg\in\widetilde{B}_{n+1}^{\prime c}, by (3), we get that hB~n+1ch\in\widetilde{B}_{n+1}^{\prime c}, so, by (4), hB~n+1ch\in\widetilde{B}_{n+1}^{c}. Thus, hA~nB~n+1ch\in\widetilde{A}_{n}\cap\widetilde{B}_{n+1}^{c} which is contained in S~\widetilde{S} by (1). Hence, by (2), we conclude that gS~g\in\widetilde{S}.

()(\subseteq) Suppose for a contradiction that there is some gS~S~g\in\widetilde{S}\setminus\widetilde{S}^{\prime}. Using (1) and (2), let n<Nn<N be maximal for which there is hA~nB~n+1ch\in\widetilde{A}_{n}\cap\widetilde{B}_{n+1}^{c} such that gker(f)hg\in\ker(f)*h. Then hS~S~h\in\widetilde{S}\setminus\widetilde{S}^{\prime} by (1) and (3). So hA~nB~n+1ch\notin\widetilde{A}_{n}^{\prime}\cap\widetilde{B}_{n+1}^{\prime c}. This together with hA~nh\in\widetilde{A}_{n} and (4) implies that hB~n+1h\in\widetilde{B}_{n+1}^{\prime}, so hker(f)hh\in\ker(f)*h^{\prime} for some hB~n+1h^{\prime}\in\widetilde{B}_{n+1}. Then gker(f)ker(f)h=ker(f)hg\in\ker(f)*\ker(f)*h^{\prime}=\ker(f)*h^{\prime}, so g=vhg=vh^{\prime} for some vker(f)=𝒥v\in\ker(f)=\mathcal{J}. Choose v𝒥v^{\prime}\in\mathcal{J} with hvh^{\prime}\in v^{\prime}\mathcal{M}. Then vg=vvh=vh=hv^{\prime}g=v^{\prime}vh^{\prime}=v^{\prime}h^{\prime}=h^{\prime}, and so hker(f)gh^{\prime}\in\ker(f)*g which is contained in S~\widetilde{S} by (2) (as gS~g\in\widetilde{S}). Therefore, by (1), hA~mB~m+1ch^{\prime}\in\widetilde{A}_{m}\cap\widetilde{B}_{m+1}^{c} for some m<Nm<N. Since hB~n+1h^{\prime}\in\widetilde{B}_{n+1}, we get m+1>n+1m+1>n+1. As gker(f)hg\in\ker(f)*h^{\prime}, we get a contradiction with the maximality of nn.

(6) follows as ff is continuous by Fact 4.10 (this is the only place in the proof where we use the assumption that uu\mathcal{M} is Hausdorff) and A~n,B~n\widetilde{A}_{n},\widetilde{B}_{n} are closed.

(7) To show the first equality, it is enough to prove that A~n\widetilde{A}_{n}^{\prime} and B~n\widetilde{B}_{n}^{\prime} are unions of fibers of ff. Let us prove it for A~n\widetilde{A}_{n}^{\prime}; the case of B~n\widetilde{B}_{n}^{\prime} is the same. So consider any pA~np\in\widetilde{A}_{n}^{\prime} and qu¯q\in\overline{u\mathcal{M}} with f(p)=f(q)f(p)=f(q). We have pvpp\in v_{p}\mathcal{M} and qvqq\in v_{q}\mathcal{M} for some vp,vq𝒥v_{p},v_{q}\in\mathcal{J}. Then f(vqp)=uvqp=up=f(p)=f(q)f(v_{q}p)=uv_{q}p=up=f(p)=f(q) and vqp,qvqv_{q}p,q\in v_{q}\mathcal{M}. So, since f|vqf|_{v_{q}\mathcal{M}} is an isomorphism, we get that q=vqpker(f)pker(f)A~n=A~nq=v_{q}p\in\ker(f)*p\subseteq\ker(f)*\widetilde{A}_{n}^{\prime}=\widetilde{A}_{n}^{\prime} by (3).

The second equality follows since:

f[A~n]=f[ker(f)A~n]=uker(f)A~n=uA~n=f[A~n],f[B~n]=f[ker(f)B~n]=uker(f)B~n=uB~n=f[B~n].\begin{array}[]{ll}f[\widetilde{A}_{n}^{\prime}]=f[\ker(f)*\widetilde{A}_{n}]=u*\ker(f)*\widetilde{A}_{n}=u*\widetilde{A}_{n}=f[\widetilde{A}_{n}],\\ f[\widetilde{B}_{n}^{\prime}]=f[\ker(f)*\widetilde{B}_{n}]=u*\ker(f)*\widetilde{B}_{n}=u*\widetilde{B}_{n}=f[\widetilde{B}_{n}].\end{array}

Hence, f[S~]f[\widetilde{S}^{\prime}] is constructible by (6). ∎

By item (5) of the claim, we get [φ(x)]u=f[f1[[φ(x)]]]=f[S~]=f[S~][\varphi(x)]\cap u\mathcal{M}=f[f^{-1}[[\varphi(x)]]]=f[\widetilde{S}]=f[\widetilde{S}^{\prime}], which is a constructible set by item (7) of the claim. ∎

Proposition 4.15.

Assume that the τ\tau-topology on uu\mathcal{M} is Hausdorff. By Corollary 4.9, uu\mathcal{M} is a compact topological group, so we have the unique normalized Haar measure huh_{u\mathcal{M}} on Borel subsets of uu\mathcal{M}. By Lemma 4.14, the formula

μu(φ(x)):=hu([φ(x)]u)\mu_{u\mathcal{M}}(\varphi(x)):=h_{u\mathcal{M}}([\varphi(x)]\cap u\mathcal{M})

yields a well-defined Keisler measure in 𝔐x(𝒢)\mathfrak{M}_{x}(\mathcal{G}) which is concentrated on u¯\overline{u\mathcal{M}}\subseteq\mathcal{M} (i.e. with the support contained in u¯\overline{u\mathcal{M}}).

Using this, the material from Lemma 6.9 to Corollary 6.12 of [CG23] goes through word for word. In particular, we get the following lemma and the main theorem describing a minimal left ideal of the semigroup (𝔐xfs(𝒢,G),)\left(\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G),\ast\right), under a more natural assumption than the property CIG1 (requiring that uu\mathcal{M} is compact with respect to the induced topology instead of the τ\tau-topology) assumed in [CG23].

Lemma 4.16.

Assume that the τ\tau-topology on uu\mathcal{M} is Hausdorff. Then μuδp=μu\mu_{u\mathcal{M}}*\delta_{p}=\mu_{u\mathcal{M}} for all pp\in\mathcal{M}, where δp\delta_{p} is the Dirac measure at pp.

Let 𝔐():={μ𝔐x(𝒢):S(μ)}\mathfrak{M}(\mathcal{M}):=\{\mu\in\mathfrak{M}_{x}(\mathcal{G}):S(\mu)\subseteq\mathcal{M}\}.

Theorem 4.17.

Assume that the τ\tau-topology on uu\mathcal{M} is Hausdorff. Then 𝔐()μu\mathfrak{M}(\mathcal{M})*\mu_{u\mathcal{M}} is a minimal left ideal of 𝔐xfs(𝒢,G)\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G), and μu\mu_{u\mathcal{M}} is an idempotent which belongs to 𝔐()μu\mathfrak{M}(\mathcal{M})*\mu_{u\mathcal{M}}.

In the next section we will see that the assumption that the τ\tau-topology on uu\mathcal{M} is Hausdorff is always satisfied when GG is a countable NIP group.

5. Revised Newelski’s conjecture for countable NIP groups

The goal of this section is to prove the revised Newelski’s conjecture (see [KP23, Conjecture 5.3]) working over a countable model. We consider here a standard context for this conjecture (originally from [New09, Section 4], but see also [CPS14]) which is slightly more general than in the previous Section 4. Let MM be a model of a NIP theory, GG a 0-definable group in MM, and NMN\succ M an |M|+|M|^{+}-saturated elementary extension. By SG,ext(M)S_{G,\textrm{ext}}(M) we denote the space of all complete external types over MM concentrated on GG, i.e. the space of ultrafilters of externally definable subsets of GG. It is a GG-flow with the action given by left translation, which is naturally isomorphic as a GG-flow to the GG-flow SGfs(N,M)S_{G}^{\operatorname{fs}}(N,M) of all complete types over NN concentrated on GG and finitely satisfiable in MM. Note that the previous context Sxfs(𝒢,G)S_{x}^{\operatorname{fs}}(\mathcal{G},G) of Section 4 is a special case when M=GM=G and N=𝒢N=\mathcal{G}. On SGfs(N,M)S_{G}^{\operatorname{fs}}(N,M) we have the left continuous semigroup operation defined in the same way as for Sxfs(𝒢,G)S_{x}^{\operatorname{fs}}(\mathcal{G},G) (i.e. pq:=tp(ab/N)p*q:=\operatorname{tp}(ab/N) for any apa\models p, bqb\models q such that tp(a/N,b)\operatorname{tp}(a/N,b) is finitely satisfiable in MM), and Fact 4.11 still holds for it.

The following revised Newelski’s conjecture was stated in [KP23, Conjecture 5.3].

Conjecture 5.1.

Assume that TT is NIP. Let \mathcal{M} be a minimal left ideal of SGfs(N,M)S_{G}^{\operatorname{fs}}(N,M) and uu\in\mathcal{M} an idempotent. Then the τ\tau-topology on uu\mathcal{M} is Hausdorff.

A background around this conjecture, including an explanation that it is a weakening of Newelski’s conjecture, is given in the introduction, and in more details in the long paragraph preceding Conjecture 5.3 in [KP23] and short paragraph following it.

In order to prove Conjecture 5.1 for countable MM, first we will deduce from the main theorem of [Gla18] on the structure of tame, metrizable, minimal flows that each such flow has Hausdorff ideal group. Then Conjecture 5.1 for countable MM will follow using this fact and a presentation of SGfs(N,M)S_{G}^{\operatorname{fs}}(N,M) as an inverse limit of certain metrizable flows. The topological dynamical material below is rather standard, but it requires recalling quite a few notions and basic facts about them, and making some observations.

From now on, until we say otherwise, we are in the general abstract context of GG-flows and homomorphisms between them, where GG is an arbitrary abstract (not necessarily definable in a NIP theory) group. We let (G,X)(G,X), (G,Y)(G,Y), etc. be GG-flows, which we will sometimes denote simply as X,YX,Y, etc.

For a proof of the following fact see [Rze18, Proposition 5.41].

Fact 5.2.

Let π:XY\pi\colon X\to Y be an epimorphism of GG-flows. Then π~:E(X)E(Y)\tilde{\pi}\colon E(X)\to E(Y) given by π~(η)(π(x))=π(η(x))\tilde{\pi}(\eta)(\pi(x))=\pi(\eta(x)) is a well-defined semigroup and GG-flow epimorphism. If \mathcal{M} is a minimal left ideal of E(X)E(X) and uu\in\mathcal{M} an idempotent, then π~[]\tilde{\pi}[\mathcal{M}] is a minimal left ideal of E(Y)E(Y) and π~(u)\tilde{\pi}(u) is an idempotent in π~[]\tilde{\pi}[\mathcal{M}]. Moreover, π~|u:uπ~(u)π~[]\tilde{\pi}|_{u\mathcal{M}}\colon u\mathcal{M}\to\tilde{\pi}(u)\tilde{\pi}[\mathcal{M}] is a group epimorphism and topological quotient map with respect to the τ\tau-topologies.

Remark 5.3.

If (G,X)(G,X) and (G,Y)(G,Y) are flows for which there exists a semigroup and flow epimorphism Φ:E(X)E(Y)\Phi\colon E(X)\to E(Y), then it is unique. In particular, the epimorphism π~\tilde{\pi} in Fact 5.2 does not depend on the choice of the epimorphism π\pi.

Proof.

Since Φ\Phi is a semigroup epimorphism, it satisfies Φ(idX)=idY\Phi(\operatorname{id}_{X})=\operatorname{id}_{Y}. Let gXE(X)g_{X}\in E(X) be the left translation by gg, and similarly gYE(Y)g_{Y}\in E(Y). Then Φ(gX)=Φ(gidX)=gΦ(idX)=gidY=gY\Phi(g_{X})=\Phi(g\operatorname{id}_{X})=g\Phi(\operatorname{id}_{X})=g\operatorname{id}_{Y}=g_{Y}. By definition, E(X)E(X) is the closure of {gX:gG}\{g_{X}:g\in G\}, so Φ\Phi is unique (as it is continuous). ∎

Let βG\beta G be the Stone-Čech compactification of GG and 𝒰e\mathcal{U}_{e} the principal ultrafilter at ee. As e.g. explained on page 9 of [Gla76], the GG-ambit (βG,𝒰e)(\beta G,\mathcal{U}_{e}) (where by a GG-ambit we mean a GG-flow with a distinguished point with dense orbit) is universal, and so there is a unique left continuous semigroup operation on βG\beta G extending the action of GG by left translation. (In fact, it is precisely the * operation on SG,ext(M)S_{G,\textrm{ext}}(M) for M:=GM:=G expanded by predicates for all subsets of GG). For any flow (G,X)(G,X), universality of βG\beta G also yields a unique action of the semigroup (βG,)(\beta G,*) on XX which is left-continuous and extends the action of GG, and which we will denote by \cdot. Fix a minimal left ideal \mathcal{M} of βG\beta G and an idempotent uu\in\mathcal{M}. Using this action, for any GG-flow (G,X)(G,X) with a distinguished point x0Xx_{0}\in X such that ux0=x0u\cdot x_{0}=x_{0} (note that such an x0x_{0} always exists), the Galois group of (X,x0)(X,x_{0}) is defined as:

Gal(X,x0):={pu:px0=x0}u,\operatorname{Gal}(X,x_{0}):=\{p\in u\mathcal{M}:p\cdot x_{0}=x_{0}\}\leq u\mathcal{M},

it is a τ\tau-closed subgroup of uu\mathcal{M} (see [Gla76, Page 13]). In the topological dynamics literature, this group is sometimes called the Ellis group of (X,x0)(X,x_{0}), e.g. see [Gla76, Page 13], where it is denoted 𝒢(X,x0)\mathcal{G}(X,x_{0}), for its basic properties.

There is an obvious semigroup and GG-flow epimorphism Φ:βGE(X)\Phi\colon\beta G\to E(X) given by Φ(p)(x):=px\Phi(p)(x):=p\cdot x. It is unique by Remark 5.3. As in Fact 4.11, βG\beta G is naturally isomorphic to E(βG)E(\beta G) via plpp\mapsto l_{p}, where lp(q)=pql_{p}(q)=p*q. Using this identification, for any GG-flow epimorphism f:βGXf\colon\beta G\to X, the induced map f~:βGE(X)\tilde{f}\colon\beta G\to E(X) from Fact 5.2 coincides with Φ\Phi.

Remark 5.4.

Let \mathcal{M} and uu\in\mathcal{M} be as above. Let (G,X)(G,X) be a flow and Φ:βGE(X)\Phi\colon\beta G\to E(X) the unique epimorphism defined above. Let 𝒩:=Φ[]\mathcal{N}:=\Phi[\mathcal{M}] and v:=Φ(u)v:=\Phi(u).

  1. (1)

    For every xXx\in X with ux=xu\cdot x=x, ker(Φ|u)Gal(X,x)\ker(\Phi|_{u\mathcal{M}})\subseteq\operatorname{Gal}(X,x).

  2. (2)

    For every xIm(v)x\in\operatorname{Im}(v), ux=xu\cdot x=x.

  3. (3)

    xIm(v)Gal(X,x)=ker(Φ|u)\bigcap_{x\in\operatorname{Im}(v)}\operatorname{Gal}(X,x)=\ker(\Phi|_{u\mathcal{M}}).

  4. (4)

    Φ|u:uv𝒩\Phi|_{u\mathcal{M}}\colon u\mathcal{M}\to v\mathcal{N} is a group epimorphism and topological quotient map with respect to the τ\tau-topologies.

Proof.

It is clear that 𝒩\mathcal{N} is a minimal left ideal of E(X)E(X), v𝒩v\in\mathcal{N} an idempotent, and Φ|u:uv𝒩\Phi|_{u\mathcal{M}}\colon u\mathcal{M}\to v\mathcal{N} a group epimorphism.

(1) Take pker(Φ|u)p\in\ker(\Phi|_{u\mathcal{M}}), i.e. Φ(p)=v\Phi(p)=v. Then px=Φ(p)(x)=v(x)=Φ(u)(x)=ux=xp\cdot x=\Phi(p)(x)=v(x)=\Phi(u)(x)=u\cdot x=x. Hence, pGal(X,x)p\in\operatorname{Gal}(X,x).

(2) Take xIm(v)x\in\operatorname{Im}(v), i.e. x=v(y)x=v(y) for some yXy\in X. Then ux=v(v(y))=(vv)(y)=v(y)=xu\cdot x=v(v(y))=(v\circ v)(y)=v(y)=x.

(3) The inclusion ()(\supseteq) follows from (1) and (2). For the opposite inclusion, consider any pxIm(v)Gal(X,x)p\in\bigcap_{x\in\operatorname{Im}(v)}\operatorname{Gal}(X,x). In order to show that pker(Φ|u)p\in\ker(\Phi|_{u\mathcal{M}}), it is enough to check that Φ(p)|Im(v)=idIm(v)\Phi(p)|_{\operatorname{Im}(v)}=\operatorname{id}_{\operatorname{Im}(v)} (because for any η𝒩\eta\in\mathcal{N} we have η=ηv\eta=\eta v, and so the map v𝒩Sym(Im(v))v\mathcal{N}\to\operatorname{Sym}(\operatorname{Im}(v)) given by ηη|Im(v)\eta\mapsto\eta|_{\operatorname{Im}(v)} is injective, in fact a group monomorphism, and v|Im(v)=idIm(v)v|_{\operatorname{Im}(v)}=\operatorname{id}_{\operatorname{Im}(v)} as vv is an idempotent). But this is trivial by the choice of pp: Φ(p)(v(y))=p(v(y))=v(y)\Phi(p)(v(y))=p\cdot(v(y))=v(y).

(4) In the situation when (G,X)(G,X) has a dense orbit (which is for example the case when (G,X)(G,X) is minimal), this follows from Fact 5.2, the existence of an epimorphism f:βGXf\colon\beta G\to X (as βG\beta G is a universal GG-ambit), and the observation that Φ=f~\Phi=\tilde{f} made just before Remark 5.4. In general, it follows from the straightforward generalization of Fact 5.2 stated in [KLM22, Fact 2.3]. ∎

Definition 5.5.

A GG-flow epimorphism π:XY\pi\colon X\to Y is almost 1-1 if the set X0:={xX:π1[π(x)]={x}}X_{0}:=\{x\in X:\pi^{-1}[\pi(x)]=\{x\}\} is dense in XX.

Remark 5.6.

If (G,Y)(G,Y) is minimal and π:XY\pi\colon X\to Y is almost 1-1, then (G,X)(G,X) is also minimal.

Proof.

We will show that for every xXx\in X we have X0E(X)x:={η(x):ηE(X)}X_{0}\subseteq E(X)x:=\{\eta(x):\eta\in E(X)\}. This implies that E(X)x=XE(X)x=X for every xXx\in X (because X0X_{0} is dense in XX and E(X)xE(X)x is closed in XX), which means that (G,X)(G,X) is minimal.

So fix xXx\in X, and consider any x0X0x_{0}\in X_{0}. Since (G,Y)(G,Y) is minimal, we can find τE(Y)\tau\in E(Y) such that τ(π(x))=π(x0)\tau(\pi(x))=\pi(x_{0}). Pick ηE(X)\eta\in E(X) satisfying π~(η)=τ\tilde{\pi}(\eta)=\tau. Then π(x0)=τ(π(x))=π~(η)(π(x))=π(η(x))\pi(x_{0})=\tau(\pi(x))=\tilde{\pi}(\eta)(\pi(x))=\pi(\eta(x)). Since x0X0x_{0}\in X_{0}, we conclude that η(x)=x0\eta(x)=x_{0}. ∎

Lemma 5.7.

If (G,Y)(G,Y) is minimal and π:XY\pi\colon X\to Y is almost 1-1, then the group homomorphism π~|u:uπ~(u)π~[]\tilde{\pi}|_{u\mathcal{M}}\colon u\mathcal{M}\to\tilde{\pi}(u)\tilde{\pi}[\mathcal{M}] is a topological isomorphism (in the τ\tau-topologies), where \mathcal{M} is a minimal left ideal of E(X)E(X) and uu\in\mathcal{M} an idempotent.

Proof.

By Remark 5.6, (G,X)(G,X) is minimal, so X=xX=\mathcal{M}x for every xXx\in X. Pick x0X0x_{0}\in X_{0}; then there is η0\eta_{0}\in\mathcal{M} such that x0Im(η0)x_{0}\in\operatorname{Im}(\eta_{0}). Choose an idempotent vv\in\mathcal{M} so that η0v\eta_{0}\in v\mathcal{M}. Then Im(η0)=Im(v)\operatorname{Im}(\eta_{0})=\operatorname{Im}(v), so x0Im(v)x_{0}\in\operatorname{Im}(v), and hence v(x0)=x0v(x_{0})=x_{0} by idempotence of vv.

Since the diagram

v{v\mathcal{M}}u{u\mathcal{M}}π~(v)π~[]{\tilde{\pi}(v)\tilde{\pi}[\mathcal{M}]}π~(u)π~[]{\tilde{\pi}(u)\tilde{\pi}[\mathcal{M}]}π~\scriptstyle{\tilde{\pi}}u\scriptstyle{u\circ}π~\scriptstyle{\tilde{\pi}}π~(u)\scriptstyle{\tilde{\pi}(u)\circ}

commutes (by definition of π~\tilde{\pi}) and, by Fact 4.8(6), the horizontal arrows are isomorphisms of semitopological groups, it is enough to show that π~|v:vπ~(v)π~[]\tilde{\pi}|_{v\mathcal{M}}\colon v\mathcal{M}\to\tilde{\pi}(v)\tilde{\pi}[\mathcal{M}] is an isomorphism of semitopological groups. By Fact 5.2, it is a group epimorphism and topological quotient map, so it remains to show that it is injective.

Suppose for a contradiction that ker(π~|v)\ker(\tilde{\pi}|_{v\mathcal{M}}) is non-trivial, i.e. there is ηv{v}\eta\in v\mathcal{M}\setminus\{v\} such that π~(η)=π~(v)\tilde{\pi}(\eta)=\tilde{\pi}(v). Then ηv=ηv=vv\eta v=\eta\neq v=vv, so η|Im(v)v|Im(v)\eta|_{\operatorname{Im}(v)}\neq v|_{\operatorname{Im}(v)}. On the other hand, by the first sentence of the proof, Im(v)=v(x0)\operatorname{Im}(v)=v\mathcal{M}(x_{0}). So there is ηv\eta^{\prime}\in v\mathcal{M} with ηη(x0)vη(x0)=η(x0)\eta\eta^{\prime}(x_{0})\neq v\eta^{\prime}(x_{0})=\eta^{\prime}(x_{0}), and hence (η)1ηη(x0)x0(\eta^{\prime})^{-1}\eta\eta^{\prime}(x_{0})\neq x_{0} (where (η)1(\eta^{\prime})^{-1} is the inverse of η\eta^{\prime} computed in vv\mathcal{M}). As x0X0x_{0}\in X_{0}, ηker(π~|v)\eta\in\ker(\tilde{\pi}|_{v\mathcal{M}}) and π~|v\tilde{\pi}|_{v\mathcal{M}} is a group morphism, we get π(x0)π((η)1ηη(x0))=π~((η)1ηη)(π(x0))=π~(v)(π(x0))=π(v(x0))=π(x0)\pi(x_{0})\neq\pi((\eta^{\prime})^{-1}\eta\eta^{\prime}(x_{0}))=\tilde{\pi}((\eta^{\prime})^{-1}\eta\eta^{\prime})(\pi(x_{0}))=\tilde{\pi}(v)(\pi(x_{0}))=\pi(v(x_{0}))=\pi(x_{0}) (where the last equality follows from the first paragraph) — a contradiction. ∎

A pair of points (x1,x2)(x_{1},x_{2}) of a flow (G,X)(G,X) is called proximal if there is ηE(X)\eta\in E(X) such that η(x1)=η(x2)\eta(x_{1})=\eta(x_{2}); it is called distal if x1=x2x_{1}=x_{2} or (x1,x2)(x_{1},x_{2}) is not proximal. Let PP denote the collection of all proximal pairs of points in XX. The flow (G,X)(G,X) is said to be proximal when P=X×XP=X\times X, and distal when P=ΔX:={(x,x):xX}P=\Delta_{X}:=\{(x,x):x\in X\}.

Fact 5.8.

The ideal group of every proximal flow is trivial.

Proof.

Let \mathcal{M} be a minimal left ideal of E(X)E(X) and uu\in\mathcal{M} an idempotent. From [Gla76, Chapter I, Proposition 3.2(3)], it follows that each pair of points in Im(u):=u[X]\operatorname{Im}(u):=u[X] is distal, so, by proximality, Im(u)\operatorname{Im}(u) is a singleton, say Im(u)={x0}\operatorname{Im}(u)=\{x_{0}\}. Then for any pp in uu\mathcal{M}, say p=uhp=uh with hh\in\mathcal{M}, and any xXx\in X, we have p(x)=u(h(x))=x0p(x)=u(h(x))=x_{0}. So p=up=u, hence u={u}u\mathcal{M}=\{u\}. ∎

Whenever π:XY\pi\colon X\to Y is a GG-flow epimorphism, let

Rπ:={(x1,x2)X2:π(x1)=π(x2)}.R_{\pi}:=\{(x_{1},x_{2})\in X^{2}:\pi(x_{1})=\pi(x_{2})\}.
Definition 5.9.

A GG-flow epimorphism π:XY\pi\colon X\to Y is said to be:

  1. (1)

    equicontinuous (or almost periodic) if for every ε\varepsilon which is an open neighborhood of the diagonal ΔX:={(x,x):xX}X×X\Delta_{X}:=\{(x,x):x\in X\}\subseteq X\times X, there exists a neighborhood δ\delta of ΔX\Delta_{X} such that g(δRπ)εg(\delta\cap R_{\pi})\subseteq\varepsilon for every gGg\in G;

  2. (2)

    distal if PRπ=ΔXP\cap R_{\pi}=\Delta_{X}.

It is sometimes assumed that the flows in the definition of equicontinuous extensions are minimal. On page 100 of [Gla76], Glasner defines almost periodic extensions of minimal flows in a different way. A proof that both definitions are equivalent for minimal flows can be found in [Aus88, Chapter 14, Theorem 1].

The following remark is well-known and follows from an argument on page 4 of [Gla76].

Remark 5.10.

An equicontinuous epimorphism of flows is distal.

Proposition 5.11.

If π:XY\pi\colon X\to Y is an equicontinuous epimorphism of minimal GG-flows and (G,Y)(G,Y) has a trivial ideal group, then the ideal group of (G,X)(G,X) is Hausdorff (with respect to the τ\tau-topology).

Proof.

Choose a minimal left ideal \mathcal{M} in βG\beta G, an idempotent uu\in\mathcal{M}, and an element xXx\in X with ux=xu\cdot x=x. Let Φ:βGE(X)\Phi\colon\beta G\to E(X) be the unique semigroup and GG-flow epimorphism considered before and in Remark 5.4. Put 𝒩X:=Φ[]\mathcal{N}_{X}:=\Phi[\mathcal{M}] and vX:=Φ(u)v_{X}:=\Phi(u). So vX𝒩Xv_{X}\mathcal{N}_{X} is the ideal group of (G,X)(G,X). Put y:=π(x)y:=\pi(x). Then uy=yu\cdot y=y (as uy=limgiπ(x)=π(lim(gix))=π(ux)=π(x)=yu\cdot y=\lim g_{i}\pi(x)=\pi(\lim(g_{i}x))=\pi(u\cdot x)=\pi(x)=y, where (gi)i(g_{i})_{i} is a net in GG converging to uu in βG\beta G). Let 𝒩Y:=π~[𝒩X]\mathcal{N}_{Y}:=\tilde{\pi}[\mathcal{N}_{X}] and vY:=π~(vX)v_{Y}:=\tilde{\pi}(v_{X}). So vY𝒩Yv_{Y}\mathcal{N}_{Y} is the ideal group of (G,Y)(G,Y) which is trivial by assumption.

Clearly Φ:=π~Φ:βGE(Y)\Phi^{\prime}:=\tilde{\pi}\circ\Phi\colon\beta G\to E(Y) is the unique semigroup and GG-flow epimorphism from βG\beta G to E(Y)E(Y), and Φ|u:uvY𝒩Y\Phi^{\prime}|_{u\mathcal{M}}\colon u\mathcal{M}\to v_{Y}\mathcal{N}_{Y} is a group epimorphism. Hence, as vY𝒩Yv_{Y}\mathcal{N}_{Y} is trivial, ker(Φ|u)=u\ker(\Phi^{\prime}|_{u\mathcal{M}})=u\mathcal{M}. As by Remark 5.4(1) ker(Φ|u)Gal(Y,y)\ker(\Phi^{\prime}|_{u\mathcal{M}})\subseteq\operatorname{Gal}(Y,y), we conclude that Gal(Y,y)=u\operatorname{Gal}(Y,y)=u\mathcal{M}.

Let FF be a τ\tau-closed subgroup of uu\mathcal{M}, and let

H(F):={clτ(VF):V is a τ-neighborhood of u in u}.H(F):=\bigcap\left\{\operatorname{cl}_{\tau}(V\cap F):V\textrm{ is a }\tau\textrm{-neighborhood of }u\textrm{ in }u\mathcal{M}\right\}.

As π\pi is almost periodic, [Gla76, Chapter IX, Theorem 2.1(4)] yields H(Gal(Y,y))Gal(X,x)H\left(\operatorname{Gal}(Y,y)\right)\subseteq\operatorname{Gal}(X,x), and together with the conclusion of the last paragraph this implies H(u)Gal(X,x)H(u\mathcal{M})\subseteq\operatorname{Gal}(X,x). Note that we proved it for any xXx\in X with ux=xu\cdot x=x, in particular for any xIm(vX)x\in\operatorname{Im}(v_{X}) (by Remark 5.4(2)). On the other hand, by Remark 5.4(3), xIm(vX)Gal(X,x)=ker(Φ|u)\bigcap_{x\in\operatorname{Im}(v_{X})}\operatorname{Gal}(X,x)=\ker(\Phi|_{u\mathcal{M}}). Hence, H(u)ker(Φ|u)H(u\mathcal{M})\subseteq\ker(\Phi|_{u\mathcal{M}}).

By Remark 5.4(4), ker(Φ|u)\ker(\Phi|_{u\mathcal{M}}) is τ\tau-closed. So, by [Gla76, Chapter IX, Theorem 1.9(3)] and the conclusion of the last paragraph, we get that u/ker(Φ|u)u\mathcal{M}/\ker(\Phi|_{u\mathcal{M}}) is Hausdorff with the quotient topology of the τ\tau-topology on uu\mathcal{M}. Since by Remark 5.4(4) we know that u/ker(Φ|u)vX𝒩Xu\mathcal{M}/\ker(\Phi|_{u\mathcal{M}})\cong v_{X}\mathcal{N}_{X} as semitopological groups, we conclude that vX𝒩Xv_{X}\mathcal{N}_{X} is Hausdorff. ∎

There are many equivalent definitions of tame flows (see Theorems 2.4, 3.2 and Definition 3.1 in [GM18]). We give the one which immediately points to a strong connection with the NIP property in model theory.

A sequence (fn)n<ω(f_{n})_{n<\omega} of real valued functions on a set XX is said to be independent if there exist real numbers a<ba<b such that

nPfn1[(,a)]nQfn1[(b,)]\bigcap_{n\in P}f_{n}^{-1}[(-\infty,a)]\cap\bigcap_{n\in Q}f_{n}^{-1}[(b,\infty)]\neq\emptyset

for all finite disjoint subsets P,QP,Q of ω\omega.

Definition 5.12.

Let (G,X)(G,X) be a flow. A function fC(X)f\in C(X) (i.e. a continuous real valued function on XX) is tame if the the family of translates {gf:gG}\{gf:g\in G\} does not contain an infinite independent sequence (where (gf)(x):=f(g1x)(gf)(x):=f(g^{-1}x)). The flow (G,X)(G,X) is tame if all functions in C(X)C(X) are tame.

The following fact is a part of the information contained in the main theorem (Theorem 5.3) of [Gla18] on the structure of tame, metrizable, minimal flows.

Fact 5.13.

Let (G,X)(G,X) be a tame, metrizable, minimal flow. Then there exists the following commutative diagram of GG-flow epimorphisms

X~{\tilde{X}}X{X^{*}}X{X}Z{Z}Y{Y}Y,{Y^{*},}η\scriptstyle{\eta}π\scriptstyle{\pi}θ\scriptstyle{\theta^{*}}ι\scriptstyle{\iota}σ\scriptstyle{\sigma}θ\scriptstyle{\theta}

where:

  1. (1)

    (G,X~)(G,\tilde{X}) is minimal;

  2. (2)

    (G,Y)(G,Y) is proximal;

  3. (3)

    θ,θ,ι\theta,\theta^{*},\iota are almost 1-1;

  4. (4)

    σ\sigma is equicontinuous.

We discussed all of the notions and facts above in order to deduce the following corollary.

Corollary 5.14.

The τ\tau-topology on the ideal group of any tame, metrizable, minimal flow is Hausdorff.

Proof.

We will be referring to items (1)–(4) in Fact 5.13. By (1), (3) and Remark 5.6, (G,X)(G,X^{*}) is minimal, and so are (G,Z)(G,Z), (G,Y)(G,Y^{*}), and (G,Y)(G,Y) as homomorphic images of (G,X)(G,X^{*}). By (2) and Fact 5.8, the ideal group of (G,Y)(G,Y) is trivial, and so is the ideal group of (G,Y)(G,Y^{*}) by (3) and Lemma 5.7. Hence, using (4) and Proposition 5.11, we get that the ideal group of (G,Z)(G,Z) is Hausdorff, and so is the ideal group of (G,X)(G,X^{*}) by (3) and Lemma 5.7. Therefore, the ideal groups of (G,X~)(G,\tilde{X}) and (G,X)(G,X) are both Hausdorff by Fact 5.2, because they are quotients of a compact topological group (namely, the ideal group of (G,X)(G,X^{*})) by closed subgroups. ∎

To apply this general corollary to our model-theoretic context, we need one more general observation, namely Lemma 5.16. To prove it, we have to recall a description of the τ\tau-closure that was stated as Lemma 3.11 in the first arXiv version of [KLM22] (the relevant section of [KLM22] was removed in the published version).

Fact 5.15.

Let (G,X)(G,X) be a flow, \mathcal{M} a minimal left ideal of E(X)E(X), and uu\in\mathcal{M} an idempotent. Then for every AuA\subseteq u\mathcal{M}, the τ\tau-closure clτ(A)\operatorname{cl}_{\tau}(A) can be described as the set of all limits contained in uu\mathcal{M} of nets (ηiai)i(\eta_{i}a_{i})_{i} such that ηi\eta_{i}\in\mathcal{M}, aiAa_{i}\in A, and limiηi=u\lim_{i}\eta_{i}=u.

Proof.

Consider aclτ(A)a\in\operatorname{cl}_{\tau}(A). Then, by the definition of the τ\tau-topology, there are nets (gi)iG(g_{i})_{i}\subseteq G and (ai)iA(a_{i})_{i}\subseteq A such that limigi=u\lim_{i}g_{i}=u and limigiai=a\lim_{i}g_{i}a_{i}=a. Note that uai=aiua_{i}=a_{i}, as aiAua_{i}\in A\subseteq u\mathcal{M}. Put ηi:=giu\eta_{i}:=g_{i}u\in\mathcal{M} for all ii. By left continuity, we have that limiηi=limigiu=(limigi)u=uu=u\lim_{i}\eta_{i}=\lim_{i}g_{i}u=(\lim_{i}g_{i})u=uu=u. Furthermore, limiηiai=limigiuai=limigiai=a\lim_{i}\eta_{i}a_{i}=\lim_{i}g_{i}ua_{i}=\lim_{i}g_{i}a_{i}=a.

Conversely, consider any aua\in u\mathcal{M} for which there are nets (ηi)i(\eta_{i})_{i}\subseteq\mathcal{M} and (ai)iA(a_{i})_{i}\subseteq A such that limiηi=u\lim_{i}\eta_{i}=u and limiηiai=a\lim_{i}\eta_{i}a_{i}=a. Since each ηi\eta_{i} can be approximated by elements of GG and the semigroup operation is left continuous, one can find a subnet (aj)j(a_{j}^{\prime})_{j} of (ai)i(a_{i})_{i} and a net (gj)jG(g_{j})_{j}\subseteq G such that limjgj=u\lim_{j}g_{j}=u and limjgjaj=a\lim_{j}g_{j}a_{j}^{\prime}=a, which means that aclτ(A)a\in\operatorname{cl}_{\tau}(A). ∎

Lemma 5.16.

Let (G,X)(G,X) be a flow, \mathcal{M} a minimal left ideal in E(X)E(X). Then there exists a minimal left ideal 𝒩\mathcal{N} of E()E(\mathcal{M}) and a semigroup and GG-flow isomorphism from \mathcal{M} to 𝒩\mathcal{N}. In particular, the ideal groups of the GG-flows XX and \mathcal{M} are isomorphic as semitopological groups (with the τ\tau-topologies).

Proof.

Denote the semigroup operation on E(X)E(X) by * and on E()E(\mathcal{M}) by \circ (although both are compositions of functions, but on different sets). Let f:E()f\colon\mathcal{M}\to E(\mathcal{M}) be given by f(p):=lpf(p):=l_{p}, where lp(q):=pql_{p}(q):=p*q. It is easy to check that ff is a semigroup and GG-flow monomorphism. Put 𝒩:=f[]E()\mathcal{N}:=f[\mathcal{M}]\subseteq E(\mathcal{M}). Thus, f:𝒩f\colon\mathcal{M}\to\mathcal{N} is a semigroup and GG-flow isomorphism. We need to check that 𝒩\mathcal{N} is a minimal left ideal in E()E(\mathcal{M}). For that, first note that \mathcal{M} is a minimal left ideal in itself. Indeed, if \mathcal{M}^{\prime} is a left ideal in \mathcal{M}, then for any pp\in\mathcal{M}^{\prime} and qE(X)q\in E(X), taking an idempotent uu\in\mathcal{M} such that pup\in u\mathcal{M}, we have p=upp=u*p, and so qp=q(up)=(qu)pq*p=q*(u*p)=(q*u)*p\in\mathcal{M}^{\prime} as quq*u\in\mathcal{M} (\mathcal{M} is a left ideal in E(X)E(X)). Hence, \mathcal{M}^{\prime} is a left ideal in E(X)E(X) which is contained in \mathcal{M}, so it must be equal to \mathcal{M} by minimality of \mathcal{M}. Since \mathcal{M} is a minimal left ideal in itself, we get that

  1. (1)

    𝒩\mathcal{N} is a minimal left ideal in itself.

On the other hand,

  1. (2)

    𝒩\mathcal{N} is a left ideal in E()E(\mathcal{M}).

To prove (2), consider any ηE()\eta\in E(\mathcal{M}) and pp\in\mathcal{M}, and we need to show that ηlp𝒩\eta\circ l_{p}\in\mathcal{N}. Using limits of nets, we easily see that there is η~E(E(X))\tilde{\eta}\in E(E(X)) such that η~|=η\tilde{\eta}|_{\mathcal{M}}=\eta. For any qq\in\mathcal{M} we have: (ηlp)(q)=η(lp(q))=η(pq)=η~(pq)=η(pq)=(ηp)q=lηp(q)=f(ηp)(q)(\eta\circ l_{p})(q)=\eta(l_{p}(q))=\eta(p*q)=\tilde{\eta}(p*q)=\eta^{\prime}*(p*q)=(\eta^{\prime}*p)*q=l_{\eta^{\prime}*p}(q)=f(\eta^{\prime}*p)(q), where ηE(X)\eta^{\prime}\in E(X) satisfies η~(τ)=ητ\tilde{\eta}(\tau)=\eta^{\prime}*\tau for all τE(X)\tau\in E(X) (such η\eta^{\prime} exists by Fact 4.4). Hence, ηlp=f(ηp)𝒩\eta\circ l_{p}=f(\eta^{\prime}*p)\in\mathcal{N} (and ηp\eta^{\prime}*p\in\mathcal{M} as \mathcal{M} is a left ideal).

By (1) and (2), we get that 𝒩\mathcal{N} is a minimal left ideal of E()E(\mathcal{M}).

Thus, f(u)𝒩f(u)\mathcal{N} is the ideal group of \mathcal{M}, and f|u:uf(u)𝒩f|_{u\mathcal{M}}\colon u\mathcal{M}\to f(u)\mathcal{N} is a group isomorphism, where uu\in\mathcal{M} is an idempotent. This isomorphism is topological (with respect to the τ\tau-topologies) by Fact 5.15 (which expresses the τ\tau-closure in terms of the semigroup operation and convergence within the minimal left ideal in question) and the above observation that f:𝒩f\colon\mathcal{M}\to\mathcal{N} is an isomorphism of left topological semigroups. ∎

We can finally prove Conjecture 5.1 for countable MM.

Theorem 5.17.

Let MM be a countable model of a theory with NIP and NMN\succ M be 1\aleph_{1}-saturated. Let \mathcal{M} be a minimal left ideal of SGfs(N,M)S_{G}^{\operatorname{fs}}(N,M) and uu\in\mathcal{M} an idempotent. Then the τ\tau-topology on uu\mathcal{M} is Hausdorff.

Proof.

Let MextM^{\textrm{ext}} be the Shelah expansion of MM obtained by adding predicates for all externally definable subsets of MnM^{n} for all n<ωn<\omega. By Shelah’s theorem [She09] (see also [CS13]), we know that Th(Mext)\operatorname{{Th}}(M^{\textrm{ext}}) has quantifier elimination, NIP, and all types in S(Mext)S(M^{\textrm{ext}}) are definable (i.e. all externally definable subsets of MextM^{\textrm{ext}} are definable). It follows that the Boolean algebra of externally definable subset of GG with respect to the original language coincides with the Boolean algebra of definable subsets of GG in the sense of the expanded language. Hence, SG,ext(M)=SG(Mext)S_{G,\textrm{ext}}(M)=S_{G}(M^{\textrm{ext}}). Thus, without loss of generality, we may assume that MM is a countable model of an NIP theory such that all types in S(M)S(M) are definable. Then SG,ext(M)=SG(M)S_{G,\textrm{ext}}(M)=S_{G}(M), and the semigroup operation on SG(M)S_{G}(M) is given by pq=tp(ab/M)p*q=\operatorname{tp}(ab/M), where apa\models p, bqb\models q, and tp(a/M,b)\operatorname{tp}(a/M,b) is finitely satisfiable in MM.

It is well-known, and observed first time in the introduction of [CS18], that NIP implies that (G,SG(M))(G,S_{G}(M)) is a tame flow. (A standard way to see it is to note that, by NIP, the characteristic functions of all the clopens in SG(M)S_{G}(M) are tame (in the sense of Definition 5.12) and separate points, and so, by Stone-Weierstrass theorem, they generate a dense subalgebra of C(SG(M))C(S_{G}(M)); then use the fact that tame functions on SG(M)S_{G}(M) form a closed subalgebra of C(SG(M))C(S_{G}(M)) to conclude that all functions in C(SG(M))C(S_{G}(M)) are tame.) However, (G,SG(M))(G,S_{G}(M)) is neither metrizable (even when the original language of MM was countable, the expanded language of MextM^{\textrm{ext}} is usually uncountable) nor minimal, so we cannot apply Corollary 5.14 directly to (G,SG(M))(G,S_{G}(M)).

Let Δ\Delta range over all finite collections of definable subsets of GG. For any such Δ\Delta, let G(Δ)\mathcal{B}_{G}(\Delta) be the Boolean GG-algebra (so closed under left translations by the elements of GG) of subsets of GG generated by Δ\Delta, and denote by SG,Δ(M)S_{G,\Delta}(M) the space of all ultrafilters of G(Δ)\mathcal{B}_{G}(\Delta). Note that SG,Δ(M)S_{G,\Delta}(M) is naturally a GG-flow (the action is by left translations), and SG(M)limΔSG,Δ(M)S_{G}(M)\cong\varprojlim_{\Delta}S_{G,\Delta}(M) as GG-flows. Let \mathcal{M} be any minimal left ideal (and so minimal subflow) of SG()S_{G}(\mathcal{M}). Let ΔSG,Δ()\mathcal{M}_{\Delta}\subseteq S_{G,\Delta}(\mathcal{M}) be the image of \mathcal{M} under the restriction map. It is clearly a minimal subflow of SG,Δ(M)S_{G,\Delta}(M), and the above isomorphism induces a GG-flow isomorphism limΔΔ\mathcal{M}\cong\varprojlim_{\Delta}\mathcal{M}_{\Delta} (see [Rze18, Lemma 6.42]).

Since (G,SG(M))(G,S_{G}(M)) is tame, so is (G,SG,Δ(M))(G,S_{G,\Delta}(M)) as a quotient of SG(M)S_{G}(M), and so is (G,Δ)(G,\mathcal{M}_{\Delta}) as a subflow of SG,Δ(M)S_{G,\Delta}(M) (using Definition 5.12 and Tietze’s extension theorem, or see e.g. [KR20, Fact 4.20]). Moreover, SG,Δ(M)S_{G,\Delta}(M) is metrizable since G(Δ)\mathcal{B}_{G}(\Delta) is countable by finiteness of Δ\Delta and countability of GMG\subseteq M (and this is the only place where we the assumption that MM is countable). Hence, Δ\mathcal{M}_{\Delta} is metrizable.

Summarizing, (G,Δ)(G,\mathcal{M}_{\Delta}) is a tame, metrizable, minimal flow, and hence its ideal group is Hausdorff by Corollary 5.14.

On the other hand, since limΔΔ\mathcal{M}\cong\varprojlim_{\Delta}\mathcal{M}_{\Delta}, [Rze18, Lemma 6.42] implies that the ideal group of \mathcal{M} equipped with the τ\tau-topology is topologically isomorphic to an inverse limit of the ideal groups (with the τ\tau-topologies) of the flows Δ\mathcal{M}_{\Delta}.

By the last two paragraphs, we conclude that the ideal group of \mathcal{M} is Hausdorff.

Finally, by Fact 4.11, E(SG(M))SG(M)E(S_{G}(M))\cong S_{G}(M) as semigroups and GG-flows, so \mathcal{M} can be identified as a GG-flow with a minimal left ideal of E(SG(M))E(S_{G}(M)). Then, by Lemma 5.16, the ideal groups of SG(M)S_{G}(M) and \mathcal{M} are topologically isomorphic. Therefore, by the last paragraph, the ideal group of SG(M)S_{G}(M) is Hausdorff. ∎

Combining Theorem 4.17 with Theorem 5.17 we thus get (using the notation in Theorem 4.17):

Corollary 5.18.

Assume that GG is countable and Th(G)\operatorname{{Th}}(G) is NIP, and let \mathcal{M} be a minimal left ideal in (Sxfs(𝒢,G),)(S_{x}^{\operatorname{fs}}(\mathcal{G},G),*) and uu\in\mathcal{M} an idempotent. Then 𝔐()μu\mathfrak{M}(\mathcal{M})*\mu_{u\mathcal{M}} is a minimal left ideal of (𝔐xfs(𝒢,G),)\left(\mathfrak{M}_{x}^{\operatorname{fs}}(\mathcal{G},G),\ast\right), and μu\mu_{u\mathcal{M}} is an idempotent which belongs to 𝔐()μu\mathfrak{M}(\mathcal{M})*\mu_{u\mathcal{M}}.

6. Acknowledgements

We thank Aaron Anderson, Martin Hils, Anand Pillay, Sergei Starchenko, Atticus Stonestrom and Mariana Vicaria for helpful conversations. Chernikov was partially supported by the NSF CAREER grant DMS-1651321 and by the NSF Research Grant DMS-2246598. Krupiński was supported by the Narodowe Centrum Nauki grant no. 2016/22/E/ST1/00450.

References

  • [ACP14] Hans Adler, Enrique Casanovas and Anand Pillay “Generic stability and stability” In The Journal of Symbolic Logic JSTOR, 2014, pp. 179–185
  • [Adl07] Hans Adler “Strong theories, burden, and weight” In preprint 2008, 2007
  • [AGK19] Uri Andrews, Isaac Goldbring and H Jerome Keisler “Independence in randomizations” In Journal of Mathematical Logic 19.01 World Scientific, 2019, pp. 1950005
  • [And23] Aaron Anderson “Generically Stable Measures and Distal Regularity in Continuous Logic” In Preprint, arXiv:2310.06787, 2023
  • [Aus88] Joseph Auslander “Minimal Flows and Their Extensions”, Mathematics Studies 153 Amsterdam: North-Holland, 1988
  • [Bay18] Martin Bays “Geometric stability theory” In Lectures in model theory, 2018, pp. 29–58
  • [BMPW16] Thomas Blossier, Amador Martin-Pizarro and Frank O Wagner “A la recherche du tore perdu” In The Journal of Symbolic Logic 81.1 Cambridge University Press, 2016, pp. 1–31
  • [BY09] Itaï Ben Yaacov “Transfer of properties between measures and random types”, Unpublished research note, http://math.univ-lyon1.fr/~begnac/articles/MsrPrps.pdf, 2009
  • [BY10] Itaï Ben Yaacov “Stability and stable groups in continuous logic” In The Journal of Symbolic Logic 75.3 Cambridge University Press, 2010, pp. 1111–1136
  • [BYCR24] Itaï Ben Yaacov, Artem Chernikov and Nick Ramsey “Randomization and Keisler measures in NSOP1 theories” In Preprint, 2024
  • [BYK09] Itaï Ben Yaacov and Jerome H Keisler “Randomizations of models as metric structures” In Confluentes Mathematici 1.02 World Scientific, 2009, pp. 197–223
  • [BYU10] Itaï Ben Yaacov and Alexander Usvyatsov “Continuous first order logic and local stability” In Transactions of the American Mathematical Society 362.10, 2010, pp. 5213–5259
  • [Cas11] Enrique Casanovas “More on NIP and related topics (Lecture notes of the Model Theory Seminar, University of Barcelona)”, http://www.ub.edu/modeltheory/documentos/nip2.pdf, 2011
  • [CG20] Gabriel Conant and Kyle Gannon “Remarks on generic stability in independent theories” In Annals of Pure and Applied Logic 171.2 Elsevier, 2020, pp. 102736
  • [CG22] Artem Chernikov and Kyle Gannon “Definable convolution and idempotent Keisler measures” In Israel Journal of Mathematics 248.1 Springer, 2022, pp. 271–314
  • [CG23] Artem Chernikov and Kyle Gannon “Definable convolution and idempotent Keisler measures, II” In Model Theory 2.2 Mathematical Sciences Publishers, 2023, pp. 185–232
  • [CGH23] Gabriel Conant, Kyle Gannon and James Hanson “Generic stability, randomizations, and NIP formulas” In Preprint, arXiv:2308.01801, 2023
  • [CGH23a] Gabriel Conant, Kyle Gannon and James Hanson “Keisler measures in the wild” In Model Theory 2.1 Mathematical Sciences Publishers, 2023, pp. 1–67
  • [Che14] Artem Chernikov “Theories without the tree property of the second kind” In Annals of Pure and Applied Logic 165.2 Elsevier, 2014, pp. 695–723
  • [Che18] Artem Chernikov “Model theory, Keisler measures, and groups” In Bulletin of Symbolic Logic 24.3 Cambridge University Press, 2018, pp. 336–339
  • [CK12] Artem Chernikov and Itay Kaplan “Forking and dividing in NTP2 theories” In The Journal of Symbolic Logic 77.1 Cambridge University Press, 2012, pp. 1–20
  • [CKR23] Artem Chernikov, Byunghan Kim and Nicholas Ramsey “Transitivity, Lowness, and Ranks in NSOP1 Theories” In The Journal of Symbolic Logic 88.3, 2023, pp. 919–946
  • [Coh60] Paul J Cohen “On a conjecture of Littlewood and idempotent measures” In American Journal of Mathematics 82.2 JSTOR, 1960, pp. 191–212
  • [CP12] Annalisa Conversano and Anand Pillay “Connected components of definable groups and o-minimality I” In Advances in Mathematics 231.2 Elsevier, 2012, pp. 605–623
  • [CPS14] Artem Chernikov, Anand Pillay and Pierre Simon “External definability and groups in NIP theories” In Journal of the London Mathematical Society 90.1 Oxford University Press, 2014, pp. 213–240
  • [CS13] Artem Chernikov and Pierre Simon “Externally definable sets and dependent pairs” In Israel Journal of Mathematics 194.1 Springer, 2013, pp. 409–425
  • [CS18] Artem Chernikov and Pierre Simon “Definably amenable NIP groups” In J. Amer. Math. Soc. 31.3, 2018, pp. 609–641
  • [CS21] Artem Chernikov and Sergei Starchenko “Definable regularity lemmas for NIP hypergraphs” In The Quarterly Journal of Mathematics 72.4 Oxford University Press UK, 2021, pp. 1401–1433
  • [DK12] Jan Dobrowolski and Krzysztof Krupiński “On ω\omega-categorical, generically stable groups” In The Journal of Symbolic Logic 77.3 Cambridge University Press, 2012, pp. 1047–1056
  • [EKP08] Clifton Ealy, Krzysztof Krupiński and Anand Pillay “Superrosy dependent groups having finitely satisfiable generics” In Ann. Pure Appl. Logic 151.1 Elsevier BV, 2008, pp. 1–21
  • [Ell57] Robert Ellis “Locally compact transformation groups” In Duke Mathematical Journal 24.2 Duke University Press, 1957, pp. 119 –125
  • [Ell69] Robert Ellis “Lectures on topological dynamics”, Mathematics lecture note series 28 W.A. Benjamin, Inc., 1969
  • [GK15] Jakub Gismatullin and Krzysztof Krupiński “On model-theoretic connected components in some group extensions” In Journal of Mathematical Logic 15.02 World Scientific, 2015, pp. 1550009
  • [Gla07] Eli Glasner “The structure of tame minimal dynamical systems” In Ergodic Theory and Dynamical Systems 27.6 Cambridge University Press, 2007, pp. 1819–1837
  • [Gla18] Eli Glasner “The structure of tame minimal dynamical systems for general groups” In Invent. Math. 211, 2018, pp. 213–244
  • [Gla76] Shmuel Glasner “Proximal Flows”, Lecture Notes in Mathematics 517 Berlin: Springer-Verlag, 1976
  • [Gli59] Irving Glicksberg “Convolution semigroups of measures.” In Pacific Journal of Mathematics 9.1 Pacific Journal of Mathematics, 1959, pp. 51–67
  • [GM18] Eli Glasner and Michael Megrelishvili “More on tame dynamical systems” In Ergodic Theory and Dynamical Systems in their Interactions with Arithmetics and Combinatorics: CIRM Jean-Morlet Chair, Fall 2016 Springer, 2018, pp. 351–392
  • [GOU13] Darío García, Alf Onshuus and Alexander Usvyatsov “Generic stability, forking, and þ-forking” In Transactions of the American Mathematical Society 365.1, 2013, pp. 1–22
  • [GPP15] Jakub Gismatullin, Davide Penazzi and Anand Pillay “Some model theory of SL(2,)SL(2,\mathbb{R}) In Fundamenta Mathematicae 229.2 Institute of Mathematics Polish Academy of Sciences, 2015, pp. 117–128
  • [HO10] Assaf Hasson and Alf Onshuus “Stable types in rosy theories” In The Journal of Symbolic Logic 75.4 Cambridge University Press, 2010, pp. 1211–1230
  • [HP11] Ehud Hrushovski and Anand Pillay “On NIP and invariant measures” In J. Eur. Math. Soc. 13.4 European Mathematical Society Publishing House, 2011, pp. 1005–1061
  • [HP18] Mike Haskel and Anand Pillay “On maximal stable quotients of definable groups in NIP theories” In The Journal of Symbolic Logic 83.1 Cambridge University Press, 2018, pp. 117–122
  • [HPP08] Ehud Hrushovski, Ya’acov Peterzil and Anand Pillay “Groups, measures, and the NIP” In Journal of the American Mathematical Society 21.2, 2008, pp. 563–596
  • [HPS13] Ehud Hrushovski, Anand Pillay and Pierre Simon “Generically stable and smooth measures in NIP theories” In Transactions of the American Mathematical Society 365.5, 2013, pp. 2341–2366
  • [HRK19] Ehud Hrushovski and Silvain Rideau-Kikuchi “Valued fields, metastable groups” In Selecta Mathematica 25.3 Springer, 2019
  • [Hru12] Ehud Hrushovski “Stable group theory and approximate subgroups” In Journal of the American Mathematical Society 25.1, 2012, pp. 189–243
  • [Hru19] Ehud Hrushovski “Definability patterns and their symmetries”, 2019 arXiv:1911.01129 [math.LO]
  • [Hru20] Ehud Hrushovski “Beyond the Lascar Group”, 2020 arXiv:2011.12009 [math.LO]
  • [Hru90] Ehud Hrushovski “Unidimensional theories are superstable” In Annals of Pure and Applied Logic 50.2 Elsevier, 1990, pp. 117–137
  • [Iba16] Tomás Ibarlucía “The dynamical hierarchy for Roelcke precompact Polish groups” In Israel Journal of Mathematics 215.2 Springer, 2016, pp. 965–1009
  • [Kha22] Karim Khanaki “Generic stability and modes of convergence” In Preprint, arXiv:2204.03910, 2022
  • [KI40] Yukiyosi Kawada and Kiyosi Itô “On the probability distribution on a compact group. I” In Proceedings of the Physico-Mathematical Society of Japan. 3rd Series 22.12 THE PHYSICAL SOCIETY OF JAPAN, The Mathematical Society of Japan, 1940, pp. 977–998
  • [KLM22] Krzysztof Krupiński, Junguk Lee and Slavko Moconja “Ramsey theory and topological dynamics for first order theories” In Trans. Amer. Math. Soc. 375.4, 2022, pp. 2553–2596
  • [KP17] Krzysztof Krupiński and Anand Pillay “Generalized Bohr compactification and model-theoretic connected components” In Math. Proc. Cambridge 163.2 Cambridge University Press, 2017, pp. 219–249
  • [KP23] Krzysztof Krupiński and Anand Pillay “Generalized locally compact models for approximate groups”, 2023 arXiv:2310.20683 [math.LO]
  • [KPR18] Krzysztof Krupiński, Anand Pillay and Tomasz Rzepecki “Topological dynamics and the complexity of strong types” In Israel J. Math. 228, 2018, pp. 863–932
  • [KR20] Krzysztof Krupiński and Tomasz Rzepecki “Galois groups as quotients of Polish groups” In J. Math. Log. 20.03, 2020, pp. 2050018
  • [New09] Ludomir Newelski “Topological dynamics of definable group actions” In The Journal of Symbolic Logic 74.1 Cambridge University Press, 2009, pp. 50–72
  • [New12] Ludomir Newelski “Model theoretic aspects of the Ellis semigroup” In Israel Journal of Mathematics 190.1 Springer, 2012, pp. 477–507
  • [New91] Ludomir Newelski “On type definable subgroups of a stable group” In Notre Dame journal of formal logic 32.2 University of Notre Dame, 1991, pp. 173–187
  • [Pil13] Anand Pillay “Topological dynamics and definable groups” In The Journal of Symbolic Logic 78.2 Cambridge University Press, 2013, pp. 657–666
  • [Pil96] Anand Pillay “Geometric stability theory” Oxford University Press, 1996
  • [Poi01] Bruno Poizat “Stable groups” American Mathematical Soc., 2001
  • [PT11] Anand Pillay and Predrag Tanović “Generic stability, regularity, and quasiminimality. In: Models, logics, and higher-dimensional categories”, CRM proceedings & lecture notes Providence, RI: American Mathematical Society, 2011
  • [Pym62] John Pym “Idempotent measures on semigroups” In Pacific Journal of Mathematics 12.2 Mathematical Sciences Publishers, 1962, pp. 685–698
  • [Rud59] Walter Rudin “Idempotent measures on Abelian groups” In Pacific Journal of Mathematics 9.1 Pacific Journal of Mathematics, 1959, pp. 195–209
  • [Rze18] Tomasz Rzepecki “Bounded Invariant Equivalence Relations, Ph.D. thesis, Wrocław”, 2018 arXiv:1810.05113 [math.LO]
  • [She09] Saharon Shelah “Dependent first order theories, continued” In Isr. J. Math. 173, 2009, pp. 1–60
  • [Sim03] Patrick Simonetta “An example of a C-minimal group which is not abelian-by-finite” In Proceedings of the American Mathematical Society 131.12, 2003, pp. 3913–3917
  • [Sim15] Pierre Simon “A guide to NIP theories” Cambridge University Press, 2015
  • [Sim20] Pierre Simon “On Amalgamation in NTP2 Theories and Generically Simple Generics” In Notre Dame Journal of Formal Logic 61.2, 2020
  • [Sto23] Atticus Stonestrom “On non-abelian dp-minimal groups” In Preprint, arXiv:2308.04209, 2023
  • [Wag00] Frank Wagner “Simple Theories” Kluwer, Dordrecht, 2000
  • [Wag24] Frank O Wagner “dp-minimal groups” In Preprint, arXiv:2403.12111, 2024
  • [Wan22] Paul Z Wang “The group configuration theorem for generically stable types” In Preprint, arXiv:2209.03081, 2022
  • [Wen54] J. G. Wendel “Haar measure and the semigroup of measures on a compact group” In Proceedings of the American Mathematical Society 5.6 JSTOR, 1954, pp. 923–929